Vous êtes sur la page 1sur 29

Randolph, M. F. (2003). Géotechnique 53, No.

10, 847–875

Science and empiricism in pile foundation design


M . F. R A N D O L P H 

Scientific approaches to pile design have advanced enor- Les méthodes scientifiques servant à la conception des
mously in recent decades and yet, still, the most funda- piles ont fait d’énormes progrès pendant les dernières
mental aspect of pile design—that of estimating the axial décennies et pourtant l’aspect le plus fondamental de ce
capacity—relies heavily upon empirical correlations. Im- travail de conception–l’estimation de la capacité axiale des
provements have been made in identifying the processes piles–s’appuie encore lourdement sur des corrélations
that occur within the critical zone of soil immediately empiriques. Des améliorations pour identifier les proces-
surrounding the pile, but quantification of the changes in sus qui se produisent dans la zone critique de sol dans le
stress and fabric is not straightforward. This paper voisinage immédiat de la pile ont été faites, mais la
addresses the degree of confidence we can now place (a) quantification des changements de contrainte et de struc-
on the conceptual and analytical frameworks for estimat- ture n’est pas simple. Cet exposé s’interroge sur le degré
ing pile capacity, and (b) on the quantitative parameters de confiance que nous pouvons désormais accorder (a)
required to achieve a design. The discussion is restricted aux cadres de travail analytiques et conceptuels pour
to driven piles in clays and siliceous sands, with particu- l’estimation de la capacité des piles et (b) aux paramètres
lar attention given to extrapolating from design ap- quantitatifs requis pour leur conception. Cette étude se
proaches derived for closed-ended piles of relatively small limite aux piles enfoncées dans des argiles et sables
diameter to the large-diameter open-ended piles that are siliceux, et extrapole à partir des méthodes conceptuelles
used routinely in the offshore industry. From a practical dérivées pour des piles fermées de diamètre relativement
viewpoint, we need design approaches that minimise petit et des grosses piles ouvertes qui sont utilisées régu-
sensitivity to the estimated pile capacity. This may be lièrement dans l’industrie offshore. D’un point de vue
achieved partly through a greater reliance on pile load pratique, nous avons besoin de méthodes conceptuelles qui
testing, where significant advances have been made in the minimisent l’importance de la capacité estimée de la pile.
last decade, but also by adopting design approaches that On peut y arriver en partie en accordant une plus grande
are focused more on guarding against unacceptable de- fiabilité aux essais de chargement de pile qui ont fait des
formation of the complete foundation. Example applica- progrès significatifs au cours des dix dernières années,
tions in the paper are drawn both from offshore mais aussi en adoptant des méthodes conceptuelles qui
applications, where current challenges include estimating s’attachent davantage à empêcher une déformation inac-
the axial capacity of ultra-thin-walled, large-diameter ceptable de toute la fondation. Les exemples donnés dans
caissons, and from onshore applications such as bridge cet exposé sont tirés des applications offshore où les
piers and piled raft foundations, where inelastic displace- difficultés actuelles sont d’estimer la capacité axiale des
ment of the piles is not only acceptable, but often caissons de gros diamètres aux parois ultra minces ; ces
essential for efficient design. exemples sont également tirés d’applications sur terre
comme les piles de ponts et les fondations radeaux à piles
pour lesquelles un déplacement non élastique des piles est
KEYWORDS: axial capacity; dynamic testing; pile driving; pile non seulement acceptable mais également, souvent, essen-
foundations; pile groups; piled rafts tiel à la réussite de la construction.

INTRODUCTION by estimation of axial capacity, even in applications such as


This paper provides an opportunity to reflect on the consid- pile groups for buildings and bridge piers, where the critical
erable advances that have been made over the last two issue is more likely to be the magnitude of displacements
decades in the design of piles and pile groups, and to under operating conditions. Indeed, one of the recommenda-
identify those aspects of pile performance that may be tions proposed later for onshore applications is to endeavour
estimated by sound conceptual models and analysis, and to weight design criteria more towards limiting displace-
those aspects where we still need to rely on empirical ments, even for the ultimate limit state, by means of non-
correlations. In the latter case, if we are to extrapolate to linear analysis of pile group response, rather than expressing
pile geometries or soil conditions outside the current data- them solely in terms of the capacity of individual piles. By
base, we must take care to ensure that the correlations are contrast, in the offshore field, particularly where individual
consistent with our understanding of mechanics and not piles are used as anchors, axial capacity plays a necessarily
distorted by limitations in the database. dominant role in design, and here the main challenge is
Much of the design of pile foundations is still dominated extrapolation to the extreme geometries now used, including
suction-installed caissons with diameters over 5 m and wall
thicknesses as low as 0.5% of the diameter.
Manuscript received 19 March 2003; revised manuscript accepted 6 In order to limit the scope to manageable proportions, this
October 2003. paper is restricted to the following topics:
Discussion on this paper closes 1 June 2004, for further details see
p. ii.
 Centre for Offshore Foundation Systems, The University of (a) axial capacity of displacement piles (driven or jacked)
Western Australia, Crawley, Australia. in clay and sand
The Centre for Offshore Foundation Systems is established and (b) the role of pile testing, and in particular interpretation
supported under the Australian Research Council’s research centres of dynamic pile tests
programme. (c) performance of pile groups and piled rafts.

847
848 RANDOLPH
This choice is consistent with my belief that we may never pile will undergo consolidation, with decrease in water
be able to estimate axial pile capacity in many soil types content and increase in mean effective stress. Outside this
more accurately than about 30%. We therefore need to rely zone, which may extend to a few times the diameter of the
on pile tests conducted early during the construction phase pile, the radial strains are tensile during equilibration
to refine the final design (generally in terms of varying the (Randolph & Wroth, 1979). The timescale of equilibration
embedded pile length, but possibly also the diameter or will be proportional to the square of the pile diameter, d,
number of piles). Hopefully, however, results from load tests and inversely proportional to a coefficient of consolidation,
may allow adequate performance of the pile group to be ch , that reflects (a) primarily horizontal drainage, and (b)
demonstrated, allowing for inelastic pile response, even partial consolidation and partial unloading of the soil domain
though extreme loads on individual piles exceed their nom- (Fahey & Lee Goh, 1995).
inal design capacity. The final phase comprises loading of the pile, resisted by
In each of the three areas above, my aim will be to separate shaft friction along the pile shaft, and end-bearing pressure
the ‘scientific’ and ‘empirical’ components on which we rely at the pile tip. The limiting shaft friction, s , will be
for design calculations, to identify any empirical correlations determined by the local radial effective stress at failure,  r9f ,
that appear inconsistent with theoretical reasoning, and to and an interface friction angle, , according to
suggest areas where improvements may be possible, either by
new analysis or by gathering more specific data to resolve s ¼  r9f tan  (1)
current uncertainties. Each of the areas is illustrated by
practical examples based on case histories. The magnitudes of  and, particularly,  r9f will depend on
the very complex processes that occur during pile installa-
tion and subsequent consolidation of the soil close to the
pile. Partial ‘healing’ of any residual shear surfaces gener-
AXIAL CAPACITY OF DRIVEN PILES IN CLAY ated during pile installation may occur, although it is also
Overview likely that  will reduce to a residual value quite rapidly as
Any scientific approach to predicting the limiting shaft slip occurs between pile and soil.
friction that may be mobilised along the shaft of a driven The dependence of pile shaft capacity on conditions in a
pile must consider the changes that occur during installation, very narrow zone in the immediate vicinity of the pile no
equilibration of excess pore pressures, and loading of the doubt contributes to the scatter in results from pile load
pile (Fig. 1). As the pile is driven, the soil immediately tests. Even on a single site, it is common for values of shaft
adjacent to the pile will undergo severe distortion and friction, normalised by the average shear strength, su , or
changes to the fabric, with a degree of remoulding and the vertical effective stress,  v90 , to vary quite widely, emphasis-
potential formation of residual shear planes (Bond & Jar- ing the sensitivity to details of the installation process. As
dine, 1991). The soil outside the immediate vicinity of the an extreme example, in the database of pile shaft friction
pile will be displaced outwards, with a strain field that measured in nine separate tests at Pentre, Chow (1997)
resembles spherical cavity expansion ahead of the pile tip, quotes values for s /su or s = v90 that range by more than
merging to cylindrical cavity expansion along the pile shaft. 35% from the average values, with no apparent trend with
In clay with moderate to low yield stress ratio, which is the depth or other soil characteristic.
main focus here, the mean effective stress in the soil The complexity of the changes in stress and fabric in the
adjacent to the pile will gradually reduce during the cyclic soil immediately adjacent to a driven pile has limited
shearing action as the pile is driven, and the interface analytical treatment of the processes involved, and most
friction angle, , will reduce to a residual value consistent practical design still relies on correlations (O’Neill, 2001).
with the high rates of shearing and relatively low level of It is now accepted that the simple correlation parameters Æ
effective stress (Lehane & Jardine, 1994), both of which (s /su ) and  (s = v90 ) are complex functions of soil para-
moderate the degree of damage. meters—in particular the yield stress ratio and, more deba-
At the end of installation, an excess pore pressure field tably, plasticity index, sensitivity and so forth. As the
will exist around the pile, arising partly from changes in undrained strength ratio, su = v90 , is also a function of the
mean effective stress due to shearing of the soil, but yield stress ratio, correlations for shaft friction that are
primarily from increases in total stress as the soil is forced functions of both shear strength and vertical effective stress
outwards to accommodate the volume of the pile. As posi- were introduced. Originally this was in the form of the
tive excess pore pressures dissipate, pore water will flow lambda coefficient (º ¼ s =ð2su þ  v90 Þ; Vijayvergiya &
radially away from the pile, and soil immediately around the Focht, 1972), and more recently the American Petroleum
Institute (API, 1993) guidelines, based on Randolph &
Murphy (1985), have proposed estimating the shaft friction
as the larger from the following two expressions:
pffiffiffiffiffiffiffiffiffiffiffi
s ¼ 0:5 su  v90
: :
s ¼ 0:5s0u 75  90v025 (2)

In all these correlations, there appears to be an effect of


pile length, or embedment ratio, L/d, with the average
normalised shaft friction decreasing with increasing embed-
ment ratio. This has been addressed by incorporating correc-
tions for values of L/d above a certain threshold (Semple &
Rigden, 1984), or by using a power law correlation such as
that proposed by Kolk & van der Velde (1996):
(a) (b) (c)  :
: : 40 0 2
Fig. 1. Three main phases during history of driven pile: (a) s ¼ 0:55s0u 7  90v03 (3)
installation; (b) equilibration; (c) loading
L=d
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 849
It is clear, however, that correlations of the type given in dL2 peak =(EA)pile
equations (2) and (3) are entirely empirical, and coefficients K¼ (5)
˜wres
of variation mostly exceed 25%.
where (EA)pile is the cross-sectional rigidity of the pile. The
reduction factor, Rf , will also be affected to some degree by
Length effect the soil stiffness (or local displacement to peak shaft fric-
The apparent decrease in normalised shaft friction with tion) and the precise shape of the load transfer curves.
increasing embedment ratio has been attributed to two main Therefore the actual reduction should be evaluated for any
mechanisms, associated respectively with the installation and given case, by means of numerical analysis. However, to a
the loading phases. The latter has been addressed by first approximation for preliminary design calculations, the
Randolph (1983), who showed that, where the load transfer reduction factor may be expressed as
response along the shaft exhibits strain-softening, progressive  2
1
failure of a pile could lead to a significant reduction in Rf  1  (1  ) 1  p ffiffiffiffi for K . 0:25 (6)
capacity. Fig. 2 shows the displacement profile down a 2 K
typical pile, and the relative states along load transfer curves with Rf taken as (approximately) unity for smaller values
at positions A, B and C. The design chart presented by of K.
Randolph (1983) showed that, for piles where the end- The strain-softening load transfer response arises from
bearing capacity was much less than the shaft capacity, the reduction of the radial effective stress,  r9, at the pile shaft
reduction factor, Rf , defined as and, more significantly, the reduction in interface friction
Qactual angle, , to a residual value. Ring shear tests suggest that
Rf ¼ (4) the softening factor, , may lie in the range 0.5–0.8 (com-
Qrigid pared with a recommendation of 0.7 in the American
where Qactual is the actual pile capacity and Qrigid is the ideal Petroleum Institute guidelines: API, 1993), with the lower
capacity of a rigid pile (calculated as the integrated peak range possible for high-plasticity clays at moderate to large
shaft friction), could be expressed as a function of (a) the effective stress levels. Ring shear tests show that most
degree of strain softening,  ¼ residual /peak , and (b) the strain-softening occurs within relatively small displacements
relative compressibility of the pile. (10–30 mm), although it is possible that ˜wres for full-scale
The pile compressibility may be expressed conveniently as piles might be somewhat larger. For modern offshore pile
the ratio of the elastic shortening of the pile, treated as a geometries, where the L/d ratio rarely exceeds 60, typical K
free-standing column subjected to a load equivalent to the values would not exceed 5–10, giving rise to reduction
ideal shaft capacity, dL(peak )average , to the local displace- factors in the range 0.65–0.9. Progressive failure can there-
ment, ˜wres, required for degradation from peak to residual fore still lead to a significant reduction in the ideal capacity.
shaft friction. Thus the compressibility factor, K, is defined The other source of length effect is that associated with
as stress changes during installation. This has been quantified
by means of measurements from instrumented piles, particu-
larly the extensive research programme undertaken at Imper-
ial College (Jardine and co-workers Bond, Lehane and
Chow). A summary of radial stress changes measured at the
Displacement profile
end of jacked pile installation from three different clay sites
has been presented by Lehane & Jardine (1994), as shown
in Fig. 3; each value of radial stress has been normalised by
τ the local cone resistance, qc . The three sites comprise
A
heavily overconsolidated London clay, a stiff glacial till
(Cowden), and a lightly overconsolidated silty clay or clayey
silt (Bothkennar).
The measured radial stresses have been fitted by power
law curves of the form
w

τ 30
n  0.6 0.35 0.2 Bothkennar
B
25 Cowden
Distance from pile tip, h/d

London
20

w 15 Profile from
strain path method
10 (Whittle, 1992)
Decay
τ
curves
5 (d/h)n

0
C 0 0.2 0.4 0.6 0.8 1.0
Normalised radial stress, σri /qc
w
Fig. 3. Radial stress changes during jacked pile installation in
Fig. 2. Progressive failure of pile in strain-softening soil clay (after Lehane & Jardine, 1994)
850 RANDOLPH
 n
 ri d 1.0
/ (7) Open-ended pile (d/t  40) 0.30
qc h 0.9
where h is the distance from the pile tip (equivalent to 0.8 0.25
L  z, where z is the depth and L is the embedded pile 0.7
length). Deduced values of n range from 0.2 to 0.6, although 0.6
0.20

δr/rpile
δr/req
reasons for the higher values of n have been discussed by
0.5 0.15
Coop & Wroth (1990). Also shown in Fig. 3 is a prediction
0.4
from Whittle (1992), using the strain path method with soil
parameters based on the Bothkennar site. By contrast with 0.3 0.10
the measured data, the analytical prediction shows a varia- 0.2 Closed-ended Open-ended 0.05
tion in normalised radial stress only in the lower few 0.1
diameters, where it reduces from a value close to unity down
0 0
to a value of 0.5, after which it remains constant. 1 2 3 5 7 10
The divergence between the ‘science’ of the strain path Normalised final radius, r/req
method prediction and the ‘empiricism’ of the fit to field
data suggests that further study of the processes involved is Fig. 4. Radial displacement field for closed- and open-ended
required, and we must explore what facet of soil behaviour, piles
or of experimental technique, may have led to this differ-
ence. For practical application, it will also be necessary to indicated, and the thicker line for r/req greater than 3.2 is
decide how to extrapolate from the field measurements, applicable to the open-ended pile. The right-hand axis gives
which are on full-displacement, closed-ended piles, to allow the radial displacement for the open-ended pile, normalised
estimation of stress changes around partial-displacement, by the actual pile radius, rpile, rather than the equivalent
open-ended pipe piles. These aspects may be explored con- radius, req. Note that, for d/t of 40, the ‘area ratio’ (of pile
veniently through the analogy of cavity expansion. wall to the gross cross-sectional area of the pile) is r ¼ 0.1
( 4t/d), and the equivalent radius is 0.32 times the actual
radius.
Cavity expansion analogy for excess pore pressures and The assumed radial expansion for the open-ended pile
equilibration times shown schematically in Fig. 4 is such as to accommodate
The analogy of cylindrical cavity expansion to model the the full wall thickness, essentially modelling the pile as a
installation of displacement piles formed the basis of early perfect sampling tube. Support for this assumption comes,
attempts to quantify stress changes due to pile installation experimentally, from the observation that, under the dynamic
(Kirby & Esrig, 1979; Randolph et al., 1979). Subsequently, conditions of pile driving, the soil plug does indeed appear
the strain path method, pioneered by Baligh at MIT (Baligh, to progress up the pile, with only small variations in the
1985, 1986), provided more realistic and detailed predictions position of the top of the soil plug relative to the original
for the strains and stress changes in the immediate vicinity ground surface.
of the pile, particularly in respect of the zone of very high The excess pore pressures generated by pile installation
stress gradients ahead of, and behind, the pile tip and the arise from two sources: changes in mean effective stress
transition to quasi steady-state conditions (in terms of nor- during shearing and partial remoulding of the soil (which
malised stresses) along the pile shaft. Comparison of the two will give rise to positive excess pore pressures for lightly
approaches shows that, ignoring the few diameters close to overconsolidated clay, and negative pore pressures for poten-
the pile tip, the radial displacement fields are extremely tially dilatant, heavily overconsolidated clay), and increases
similar apart from immediately adjacent to the pile shaft in mean total stress due to outward ‘expansion’ of the soil
(within a zone of thickness about 10% of the pile radius, for to accommodate the pile volume.
a full-displacement pile). Simple cavity expansion theory, applied to an elastic,
Assuming that pile installation occurs under undrained perfectly plastic soil with shear modulus G and undrained
conditions, the radial displacement, r, for soil at final shear strength su would give rise to an excess pore pressure
radius, r, may be deduced as distribution of (Gibson & Anderson, 1961)
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi    
r r r 2 ˜u rG r
¼  1 (8) ¼ ln  2 ln >0 (10)
req req req su su rpile
where req is the pile radius for a closed-ended pile, and for Although this expression does not account for changes in
an open-ended pile is the radius of an equivalent solid pile mean effective stress as the soil is sheared and remoulded,
that gives the same volume of displaced soil. For thin-walled these may be accounted for approximately for lightly over-
piles of wall thickness t the equivalent pile radius and consolidated clays by adjustment of the rigidity index, Ir ¼
diameter are G/su . The main features of a logarithmic decay with radius,
pffiffiffiffiffi pffiffiffiffiffi and typical values of maximum pore pressure adjacent to
req  dt; d eq  2 dt (9)
full displacement piles of 4su to 6su (in lightly overconsoli-
where it is assumed implicitly that the pile is installed in an dated soils) agree well with results from the strain path
unplugged manner, with the top of the internal soil plug method (Baligh, 1986). An important feature of equation
remaining (approximately) level with the external soil (10) is the term accounting for the area ratio, r, for open-
surface. ended piles, where the reduction in excess pore pressure
The relationship in equation (8) is shown in Fig. 4 for a compared with a full-displacement (solid or closed-ended)
closed-ended pile (req ¼ rpile ), and also an open-ended pile pile is su ln(r).
for a d/t ratio of 40, which is a typical value for steel pipe The excess pore pressure fields around closed-ended and
piles. It is shown dashed in the region close to the (solid) open-ended piles (with d/t ¼ 40) based on cylindrical cavity
pile, where the cavity expansion solution is no longer expansion, taking G/su ¼ 100, are shown in Fig. 5, together
deemed accurate. The location of the open-ended pile is with isochrones during dissipation. Baligh (1986) has com-
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 851
5 1·0
T0
0·9
∆u G/su  100
T  0.01 ∆umax 0·8
4
0.9 T  0.1 0·7
0·6
. T  0.8
3 07 0·5
T2
∆u/su

Closed
0·4
d/t  20
T  10 0·3
2 0.5 d/t  40
0·2
d/t  80
. 0·1
1 03 d/t  160
0
0.001 0.01 0.1 1 10
∆u/∆umax  0.1 T  cht/d 2
0
1 3 5 7 9
Normalised radius, r/req 1·0
∆u
∆umax 0·9
5 G/su  100
0·8
Pile wall T0
0·7
Teq  0.03
4 0·6
Closed
Teq  0.2 0·5 d/t  20
Teq  0.8 0·4
3 d/t  40
Teq  2 0·3
∆u/su

d/t  80
∆u/∆umax  d/t  40, rpile  3·2req 0·2
0.9 Teq 10 d/t  160
2 0·1
0.7 SPM: Whittle
0
0.5 0.001 0.01 0.1 1 10
1
0.3 Teq  cht/deq2
0.1
0 Fig. 6. Dissipation of excess pore pressures at pile shaft
1 3 5 7 9
Normalised radius, r/req
ch  15 p9 k h =ªw , where ªw is the unit weight of water. The
Fig. 5. Excess pore pressures generated by pile installation: (a) figure shows that the dissipation curves for all the piles of
closed-ended pile; (b) open-ended pile different wall thickness fall in a narrow band, when ex-
pressed in terms of Teq, based on the equivalent pile
diameter, rather than the true diameter. Note also that the
mented that the excess pore pressures predicted by cavity timescale of consolidation is affected by the original magni-
expansion may be overestimated, as a result of not following tude of the excess pore pressure ratio, ˜umax /su (and hence
the correct strain path, but for lightly overconsolidated soils the lateral extent of the pore pressure field), and the results
this will be offset by ignoring the excess pore pressures due shown in Fig. 6 are for an initial excess pore pressure ratio
to shearing of the clay (with corresponding reduction in of 4.6, corresponding to G/su ¼ 100 for the cavity expansion
mean effective stress). The isochrones of excess pore pres- analogy.
sure shown in Fig. 5 have been derived using the radial In passing, it may be noted that, in their analysis of
consolidation solution of Randolph & Wroth (1979), with dissipation around a piezocone, Teh & Houlsby (1991)
the non-dimensional times expressed as proposed a ‘generalised’ time factor, T  , given by
ch t ch t 4ch t

d 2
; Teq ¼ 2
d eq
(11) T ¼ pffiffiffiffi (12)
d2 I r
where t is the time and ch is an appropriate coefficient of in order to bring together dissipation curves for different soil
consolidation for horizontal drainage. During the consolida- rigidity indices. This contrasts with the normalisation using
tion process, the outer soil (beyond 3–5 pile radii) under- Teq in Fig. 6(b), where normalisation using deq can be shown
goes swelling, while the inner soil consolidates: hence the to be equivalent to taking Teq as inversely proportional to Ir
coefficient of consolidation must reflect this fact (Fahey & (rather than to the square root of Ir ). In fact, the optimal
Lee Goh, 1995), and is most easily assessed through piezo- normalisation depends on (a) the range of Ir values that need
cone dissipation tests. to be considered, and (b) whether the focus is on the early
An interesting (and somewhat surprising) feature of Fig. 5 dissipation response (up to T50 ) or the later response (times
is that isochrones for equal proportions of excess pore greater than T50 ). For the interpretation of piezocone tests,
pressure, ˜u/˜umax , occur at very similar non-dimensional with a likely range for Ir between 50 and 500 and with the
times, T and Teq for the two pile types, as remarked on by focus on the early dissipation response, the normalisation
Whittle (1992). This is illustrated in Fig. 6, where the proposed by Teh & Houlsby (1991) may be optimal. How-
normalised excess pore pressure is plotted against the two ever, dissipation around open-ended piles, where the focus is
alternative time factors, for closed and open-ended piles of more on the times for 50–90% dissipation, and the range of
different wall thickness ratios. For comparison, a dissipation rIr (in the light of equation (10)) that need to be considered
curve based on the strain path method, as presented by is very broad, the normalisation shown in Fig. 6(b) using Teq
Whittle (1992), is also shown. Whittle’s original curve was appears more useful.
presented using a time factor expressed in terms of vertical Despite the approximations involved in the cylindrical
pre-consolidation pressure,  p9 , and horizontal permeability, cavity analogue for pile installation, it appears that
kh , and the results in Fig. 6 have been scaled by assuming the general pattern of excess pore pressure, and the
852 RANDOLPH
consolidation response, can be predicted reasonably for both that simple scaling of piezocone dissipation times, by the
piezocones and driven piles. Two examples illustrating this square of the diameter ratio (equivalent pile diameter di-
are shown in Fig. 7. The first example is for a closed-ended vided by piezocone diameter), should give a reasonable
driven pile, which was subjected to dynamic tests at differ- estimate of the consolidation times for a pile.
ent times after driving (data kindly supplied by Mr Antonio Referring to Fig. 6, two important observations may be
Alvez, PhD student at COPPE, Federal University of Rio de made. The first is that dissipation times for typical open-
Janeiro). Dissipation from a piezocone test, from which ended piles (d/t  40) and suction caissons (d/t  200) will
appropriate values of Ir and ch were deduced, is compared be respectively one and two orders of magnitude shorter
with theoretical dissipation curves and also the measured than for a closed-ended pile of the same diameter. The
increase of shaft resistance, Qs (where Qs,1 and Qs0 repre- second observation is that significant dissipation, with 20%
sent long-term and initial shaft resistance respectively). Two reduction in pore pressure, occurs for Teq  0.1. For typical
different dynamic pile–soil interaction models were used in values of consolidation coefficient in the range 3–30 m2 /yr,
analysing the pile tests, a continuum model and the Smith this value of Teq corresponds to 0.5–5 days for an open-
(1960) model, both of which are described later. Although ended offshore pile 2 m in diameter, or 0.3–3 days for a
the increase in shaft friction is not precisely proportional to closed-ended onshore pile 0.5 m in diameter. These times
the decrease in excess pore pressure, because of stress are longer than most installation times (except in the case of
relaxation effects (discussed later) and changes in radial equipment breakdown). However, for the 0.1 m diameter
effective stress during loading, it seems that consolidation instrumented pile used to obtain the data in Fig. 3, 20%
theory gives a sufficiently accurate estimate of the timescale pore pressure dissipation would occur in 0.3–3 h, compared
of increase in shaft resistance. with total jacking periods of 1–5 h (Lehane & Jardine,
The second example, in Fig. 7(b), is from centrifuge 1994).
model tests on very thin-walled suction caissons, reported by It appears, therefore, that partial pore pressure dissipation
Cao et al. (2002). The prototype dimensions of the caisson during installation may account, at least in part, for the h/d
are shown, with a d/t ratio of 80. However, as the caisson effect deduced from radial stress measurements, and the
was installed using suction, the outward soil movement may divergence between the trends in the data and theoretical
be less than for a driven pile, as more soil is drawn inside predictions from the strain path method. Rapid initial pore
the caisson (Andersen & Jostad, 2002). Hence the operative pressure dissipation may also account for the low  values
d/t ratio may be nearer 160 than 80 in terms of outward reported by Karlsrud (1999) in low-plasticity clays. Such
movement of soil. clays, with high silt content, are likely to show shorter
In both of these examples, the experimental data are consolidation times, comparable with pile installation times,
matched reasonably well by the theoretical dissipation leading to greater damage to the soil (lower residual inter-
curves. The first example, in Fig. 7 (a), also demonstrates face friction angles, because of the higher effective stress
levels during installation), less ‘set-up’ following installation,
and thus lower shaft friction values than for higher plasticity
clays.
1·0 0
∆u
∆umax
0·8 0·2

Radial stress changes during installation, equalisation and


0·6 0·4 loading
The pile shaft friction depends on the radial effective
0·4 Theory: G/su  50 0·6
Theory: G/su  100
stress acting around the shaft, according to equation (1), and
Cone: Mid-face this may be estimated by considering the sequential changes
0·2 Cone: Shoulder 0·8 during pile installation, consolidation and loading. Measure-
Qs  Qs0
Pile: Continuum
Qs  Qs0 ments of radial total stress, ri (less the in situ pore pressure,
Pile: Smith model
0 1·0 u0 ) immediately after installation, and radial effective stress,
0·001 0·01 0·1 1 10 100
 r9c , at the end of consolidation, both normalised by the in
Normalised time, T  cht/d 2 situ vertical effective stress,  v90 , are shown in Fig. 8 for
(a) values of h/d . 10. The data were assembled by Lehane
(1992), and Fig. 8(a) shows his proposed trend lines (see
also Lehane et al., 1994).
1·0
∆u During installation, the trend of radial total stress ratio,
∆umax
0·8
ð ri  u0 Þ= v90 , increases in proportion to the yield stress
65 mm ratio to the power of about 0.4, from a value of 2 for
24 m normally consolidated soil, to just under 10 at very high
0·6
yield stress ratio. As Lehane (1992) observed, the gradient is
0·4
5·2 m approximately parallel to the correlation of K0 with over-
Theory: d/t  80 consolidation ratio proposed by Mayne & Kulhawy (1982),
0·2 Theory: d/t  160 with a radial total stress ratio of 3–3.5 times K0 .
Test SAT06 Although the trend in the data on Fig. 8(a) is evident, the
Test SAT08 logarithmic scales can lead to quite significant deviation
0
0·0001 0·001 0·01 0·1 1 from the mean line. At present, the only viable analytical
Normalised time after installation , T  cht/d 2 approach for quantifying detailed stress changes during pile
(b) installation and consolidation appears to be the strain path
Fig. 7. Measured dissipation around closed and open-ended method, but simpler quasi-analytical approaches are needed
piles: (a) piezocone dissipation and pile shaft resistance in for routine design. Potential approaches, admittedly some-
high-plasticity clay (data provided by Mr Antonio Alvez); (b) what speculative, are discussed here.
pore pressure dissipation around thin-walled caisson (data from After installation, the radial total stress (less the in situ
Cao et al., 2002) pore pressure) may be expressed as
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 853
10 Such a trend is consistent with radial consolidation mod-
After installation
(σir  u0)/σv′0 els, which show the outer soil (beyond 3–5 times the pile
Radial stress coefficients

radius) swelling, while the inner core consolidates (Fahey &


(σir  u0)/σ′v0 and σ′rc/σ′v0

Lee Goh, 1995). It is the difference in stiffness of these two


zones that gives rise to the relaxation in total radial stress
(with no relaxation in classical solutions where the soil is
1
After consolidation
assumed elastic and homogeneous). The relaxation gradient
Increasing σr′c/σv′0 at any stage during consolidation is d r9=du, and it may be
sensitivity argued that this quantity will become progressively less than
unity the softer the inner soil is relative to the outer
(swelling material), and will therefore be a function of the
0·1 relative magnitude of the current radial effective stress,  r9,
1 10 100 and the preconsolidation or yield stress,  v9c . This effect may
Yield stress ratio, R be captured by a function such as
(a) d r9
 ¼ ºe( r9 r9i )= v9c (14)
du
10 where º and  are adjustable parameters.
After installation
(σir  uo)/σv′0
Integrating this expression over the change in excess pore
pressure from ˜umax down to zero, the final radial effective
Radial stress coefficients
(σir  u0)/σ′v0 and σ′rc/σ′v0

stress is given by
 
 r9c  r9i R º ˜umax
¼ þ ln 1 þ (15)
1  v90  v90  R  v90
After consolidation
Equation (15) σr′c/σv′0 where R is the yield stress ratio,  v9c = v90 . This expression is
plotted in Fig. 8(b), adjusting º to unity and  to a value of
5, in order to give a reasonable fit to the data (identical to
the data in Fig. 8(a)). Although this approach is speculative,
0·1 and significant further work is needed before it might be
1 10 100 useful in design, the concept of a relaxation gradient that
Yield stress ratio, R varies during consolidation is consistent with physical argu-
(b) ments of the conditions around the pile, and also with field
measurements by Lehane (1992), which indicate a gradual
Fig. 8. Radial stress coefficients after installation and consolida- reduction of jd r9=duj during consolidation. One important
tion (data from Chow, 1997): (a) relaxation ratio as function of consequence is that the net relaxation ratio, ( ri  u0 )= r9c ,
soil sensitivity; (b) relaxation ratio derived from function of will be higher (for a given soil) for an open-ended pile than
current yield stress ratio for a closed-ended pile.
The final phase of the pile’s history to consider is the
 ri  u0 ¼  r9i þ ˜umax ¼ ( r9i  p9i ) þ p90 þ ˜ p (13) loading phase. By the end of consolidation, the radial effec-
tive stress will have become the largest of the three normal
where ˜umax is the maximum excess pore pressure and p90 stresses (vertical, radial and circumferential) close to the
and p9i are respectively the original in situ mean effective pile. During loading of the pile, a reduction in the radial
stress and the value just after pile installation (adjacent to stress is therefore expected. Lehane (1992) and the design
the pile shaft). The bracketed term has a relatively narrow approach proposed by Jardine & Chow (1996) suggest that
range [negative, owing to the slight unloading strains next to the reduction may be taken as about 20%, independent of
the pile according to the strain path method (Baligh, 1986; the yield stress ratio, so that  r9f in equation (1) is then
Whittle, 1992), but limited in magnitude to the current 0:8 r9c .
undrained shear strength, allowing for any remoulding that
may have occurred as the pile is installed]. The in situ mean
effective stress, p90 , may be estimated through K0 , and the Example design calculations: new horizons
increase in mean total stress, ˜p, required to accommodate The offshore industry continues to face new challenges as
the pile should prove amenable to estimation through numer- it moves into deeper waters and new regions of the world.
ical analysis (strain path or cavity expansion methods). Currently, one of the most active offshore areas is off the
Estimating these quantities with any accuracy at present is west coast of Africa, where very high-plasticity clays have
not straightforward, but the approach represents a possible been encountered in water depths of 1000 m. Typical ‘gener-
scientific way forward. ic’ soil properties, based on data from a number of sites, are
During equilibration, the excess pore pressure reduces to summarised in Table 1.
zero and the radial effective stress increases to a final value The combination of high plasticity index with high fric-
denoted by  r9c. The data for the final radial effective stress tion angles measured in triaxial compression and simple
ratio,  r9c = v90 , in Fig. 8(a) have been correlated with lines shear is unusual and, in a similar fashion to Mexico City
that lie nearly parallel to the trend of the installation clay, lies well outside common correlations of friction angle
stresses, but are offset by varying amounts, depending on with PI (Mesri et al., 1975). Unlike Mexico City clay,
the sensitivity of the clay (Lehane, 1992; Jardine & Chow, however, interface friction angles are significantly lower,
1996). During consolidation there is some relaxation in total particularly at residual. This characteristic poses a particular
stress (so that the final radial effective stress is less than the challenge in estimating the shaft capacity of driven pipe
initial radial total stress), and the data suggest that the piles and thin-walled suction caissons in these clays, as
degree of relaxation is high for low yield stress ratios (also traditional approaches based on correlations with su and  v90
high sensitivity) and reduces as the yield stress ratio in- will diverge from more fundamental approaches based on
creases (and sensitivity reduces). equation (1).
854 RANDOLPH
Table 1. Clay properties offshore West Africa
Parameter Typical values
Shear strength, su : kPa .
1 5z (with z the depth in m)
Effective unit weight, ª9: kN/m3 3.5
Yield stress ratio, R 1.8
Sensitivity, St 4
Plasticity index, PI: % 100
Friction angle, 9: degrees 35
Interface friction angle, : degrees 20 (residual value 12)

To illustrate the ideas discussed earlier, two different Shaft friction: kPa
geometries of offshore piles will be considered: 0 40 80 120 160
0
(a) a conventional pipe pile, 2 m in diameter and with
50 mm wall thickness (d/t ¼ 40, r ¼ 0.1) embedded 20 Shear strength profile
100 m (L/d ¼ 50)
(b) a suction caisson, 6 m in diameter with wall thickness 40

Depth: m
of 30 mm (d/t ¼ 200, effective area ratio, allowing
for suction installation, of r ¼ 0.01) embedded 20 m 60
(L/d ¼ 3.3). MTD
80 approach
The shaft capacity of these piles will be estimated using the
approach described here (equations (10), (13) and (15)) and 100 Present
also the method of Jardine & Chow (1996) (referred to here approach API (1993)
as the MTD method, as it is known in the offshore industry). 120
(a)
The MTD method for piles in clay is based on the empirical
correlations of Lehane (1992) and Lehane et al. (1994). In
contrast to the MTD method, the effect of h/d is ignored in
the present approach apart from for h/t , 10, where the
normalised radial total stress is assumed to increase gradu- Shaft friction: kPa
ally by a maximum factor of 2 at the pile tip (as suggested 0 5 10 15 20 25 30
by the strain path method results shown in Fig. 3). Simplis- 0
tically, the radial effective stress just after installation ( r9i )
has been estimated using equation (1), assuming that the 5
Shear strength
shaft friction during installation is equal to the remoulded
shear strength, and the maximum excess pore pressure gen-
Depth: m

10 MTD
erated by a solid pile has been taken as 4.6su . approach
The profiles of peak shaft friction obtained from these 15
two approaches are compared with that estimated using the
API guidelines (API, 1993) in Fig. 9 for each pile geometry. Present
20
In Fig. 9(a), the strong h/d effect from the MTD method is approach
evident, with lower shaft friction over most of the pile shaft, API (1993)
apart from close to the tip. For the particular combination of 25
(b)
soil properties, it turns out that the approach described here
gives a shaft friction profile that is remarkably similar to Fig. 9. Profiles of peak shaft friction for offshore piles: (a)
that obtained from API (1993), although this is something of driven pipe pile (L/d 50, r 0.1); (b) suction caisson (L/d
a coincidence and will be affected by the interface friction 3.3, r 0.01)
angle, . Average values of shaft friction from the different
methods are quite close, with the MTD method about 10%
lower than the other two methods. slip of 50 mm, are shown in Fig. 10. The reduction factor
For the caisson (Fig. 9(b)), there is a much greater due to strain-softening is around 10%, which is somewhat
divergence of shaft friction profiles. The API (1993) profile less than the value of 15% estimated from equations (5) and
is identical to that for the driven pile, whereas the approach (6), mainly because of the triangular distribution of shaft
suggested here gives lower shaft friction, largely because of friction arising from the linearly increasing shear strength
the low area ratio of the caisson and hence lower excess with depth.
pore pressures generated during installation and lower final
radial effective stresses. The MTD method was not intended
to apply to piles with such low L/d or high d/t ratios, which Summary
fall well outside the database used to calibrate the method. The ‘science’ in estimating driven pile capacity in clay
It is an instructive comparison, however, reminding us that provides the framework within which the different phases of
extrapolation of any design method must be carried out with the installation, consolidation and loading history of the pile
care, particularly where the method is based on empirical are considered. It also extends to different analytical ap-
correlations. proaches, such as the strain path method, and cavity expan-
Whereas the suction caisson may be considered as effec- sion, which allow quantification of certain aspects of each
tively rigid, in terms of strain-softening effects during axial process. Magnitudes of total stress increase, quantification of
loading, the driven pile is relatively flexible. The calculated the differences between full and partial displacement piles,
load–displacement responses, assuming strain softening by and estimation of the timescale for consolidation may all be
40% ( decreasing from 208 to 128) over relative pile–soil treated analytically. However, design calculations still rely on
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 855
40 000 tionship between stiffness and stress, both contribute to a
decreasing gradient of base resistance with depth (Randolph
35 000 Ideal et al., 1994). For shaft friction, although equation (1) still
capacities
30 000 Present provides the physical basis, the normal effective stress,  r9f ,
approach at any given depth has been found to degrade as the pile is
Pile head load: kN

25 000 installed, owing to gradual densification of the surrounding


material. These components of the axial capacity of driven
20 000
piles in sand, and the necessary adjustments for open-ended
MTD method
15 000 piles, are explored here in the context of recent design
recommendations (Jardine & Chow, 1996).
10 000

5 000
Base resistance
0 Although it is natural to correlate the end-bearing resis-
0 0·02 0·04 0·06 0·08 0·10 tance of a pile with the cone resistance, consideration must
Pile head displacement: m be given to the displacement needed to mobilise a given
proportion of cone resistance. Fleming (1992) proposed a
Fig. 10. Load–displacement response of driven pipe pile hyperbolic relationship for bored piles, relating the end-
bearing pressure, qb , and the base displacement, wb, giving a
normalised end-bearing resistance, qb /qc , expressed as
empirical correlations in order to quantify those aspects that
are dominated by the complexities of soil response, such as qb wb =d
 (16)
reduction in effective stresses and degree of remoulding qc wb =d þ 0:5qc =Eb
during pile installation, relaxation of radial total stress
where Eb is the Young’s modulus of the soil below the pile
during consolidation, and reduction in radial effective stress
base.
during loading.
For a bored pile, with initially zero base pressure at zero
There appears to be divergence between analysis and field
displacement, this relationship will lead to end-bearing pres-
measurements in respect of the h/d effect during pile instal-
sures mobilised at a base displacement of 0.1d of around
lation, although partial consolidation appears partly respon-
15–20% of qc (Lee & Salgado, 1999). However, for driven
sible. Resolution of this is important, and requires careful
and jacked piles, significant residual pressures are locked in
review of what fundamental mechanisms might lead to an
at the pile base during installation (equilibrated by negative
h/d effect. An improved model to quantify stress relaxation
shear stresses along the pile shaft, as if the pile were loaded
during consolidation is also needed, perhaps through numer-
in tension). This will lead to a stiffer overall pile response in
ical parametric studies, as this is an area where considerable
compression, and significantly higher end-bearing stresses
scatter in the database exists. The concept of a relaxation
mobilised at small displacements.
gradient that changes as consolidation proceeds, as the
The magnitude of residual base stress will depend on the
relative stiffness of the inner and outer soil zones evolves,
relative magnitudes of shaft and base capacity, as well as on
has been proposed as a possible way forward.
the method of installation. For jacked piles the residual base
The challenge of providing anchors in deepwater, using
stress can be as high as 70–80% of the lesser of shaft or
suction-installed caissons, requires extension of our current
(ultimate) base capacity (Poulos, 1987). For driven closed-
design approaches to low aspect ratio (L/d , 6) and low
ended piles the residual stress will be lower, but may still be
area ratio (r  0.01) pile geometries. A rational scientific
as high as 75% of the base capacity (Maiorano et al., 1996).
basis is essential for this.
The lowest residual base stress is likely to be for open-ended
piles, unless they become fully plugged during driving.
Equation (16) can be generalised to allow for a residual
AXIAL CAPACITY OF DRIVEN PILES IN SAND pressure, qb0 , locked in below the pile base at the start of
Overview loading, to give
Over the last decade there have been two major advances
in design approaches for driven piles in sand. The first of qb wb =d þ 0:5qb0 =Eb
 (17)
these is the capturing, through instrumented pile tests, of the qc wb =d þ 0:5qc =Eb
gradual degradation of shaft friction at any given depth as The resulting end-bearing responses are illustrated in Fig. 11
the pile is driven progressively deeper (Lehane et al., 1993), for Eb /qc ¼ 1.25 (lower set of curves for each value of
and the second is the linking of key parameters such as base qbo /qc ) and Eb /qc ¼ 5 (upper set of curves for qb0 /qc ¼ 0.3
resistance and maximum shaft friction to the cone resistance, and 0.7). This range of Eb /qc reflects conservative values
qc , which has evolved from the early correlations of suggested for bored and driven piles (Poulos, 1989; Fleming,
Bustamante & Gianeselli (1982). Both of these advances are 1992).
empirical in nature, but they embody principles that could, The exact form of the end-bearing response is of course
in due course, be quantified more scientifically. debatable. However, the main principles illustrated in Fig. 11
Historically, pile design in sand has been based on simple are as follows:
linear relationships for both shaft friction and base resis-
tance, but with limiting values at some ‘critical depth’ (a) Steady-state conditions are reached after large displace-
expressed either in absolute terms or normalised by the pile ment (4–10 diameters for zero residual stress),with the
diameter (Vesic, 1967, 1970; Coyle & Castello, 1981). The end-bearing resistance of a pile approaching the cone
rationale behind this approach has been challenged resistance, after appropriate averaging of the latter
(Kulhawy, 1984), and alternative explanations offered for the quantity to reflect the larger size of the pile.
experimental finding that increasing lengths of piles driven (b) At limited displacements, such as 10% of the pile
into sand do not yield proportional increases in capacity. diameter as is often taken as the practical definition of
For base resistance, the influence of decreasing friction ‘ultimate’, the end-bearing resistance will be signifi-
angle with increasing stress level, and the non-linear rela- cantly lower than the cone resistance, and will also
856 RANDOLPH
 
1·0 qbu d
¼ 1  0:5 log > 0:13 (18)
qc d cone
0.8
At first glance, the design curve appears a reasonable fit to
Eb /qc  5
0.9 the data, in spite of some scatter. However, the data for
0.6 small pile diameters are dominated by jacked piles, where
qbo /qc  0.7
the full cone resistance (appropriately averaged according to
qb /qc

0.6
the pile diameter) would be mobilised at each stroke, and
0.4 Eb /qc  1.25 high residual stresses (or at least a high reloading stiffness)
0.3
will be retained. An annotated version of the database is
qbo/qc  0.3 shown in Fig. 13(a), with jacked piles indicated and also a
0.2 0 vibro-driven pile, where the normalised end-bearing capacity
0 0.1 0.2
falls below the other data. The driven pile result from the
Akasaka (AK: BCP Committee, 1971: see legend in Fig. 12)
0 pile tests plots above the jacked pile data, but the reported
0 1 2 3 4
load–displacement plot (see their figure 9) is anomalous,
Normalised displacement, wb /d
with a base resistance that suddenly falls after a displace-
Fig. 11. Development of end-bearing resistance ment of one pile diameter, with a corresponding jump in the
shaft friction (the total load remaining largely unchanged).
Correction for that anomaly would result in a normalised
end-bearing (qbu /qc ) of 0.4 for the driven pile.
depend strongly on any residual stresses locked in at Load cells or strain gauges in instrumented driven piles
the pile base at zero displacement. tend to undergo zero shifts during installation, due to
changes in fabrication strains within the pile caused by the
The large displacements necessary to mobilise a true ‘ulti- high dynamic stresses. It is therefore usually necessary to
mate’ end-bearing capacity lead to a form of scale effect in zero strain gauges prior to testing the pile statically, and to
comparing pile end-bearing and cone resistance, as the rely on alternative means to estimate any residual base loads.
(much greater diameter) pile will react more slowly to A common approach is to assume equal shaft capacity in
changes in stratigraphy than the cone. To overcome this, it is tension and compression, although this will tend to over-
essential to average the cone resistance over a number of estimate residual base loads (see later). Correction for
diameters above and below the pile base level. Distances residual loads varies among the pile tests, but examples of
recommended range up to 8 diameters above the pile base how this may change the deduced end-bearing capacity are
(to allow for the gradual development of residual stresses), shown in Fig. 13(a) for the Baghdad (BG: Altaee et al.,
and 2 diameters below the pile base, with a weighting 1992, 1993) and Hunters Point (HP: Briaud et al., 1989) pile
towards the minimum envelope of the cone profile (Fleming tests (the arrows linking uncorrected and corrected data).
& Thorburn, 1983). In practice, averaging over a shorter From the current database, however, and notwithstanding
length, between 1 and 2 diameters above and below the pile some inconsistency in respect of allowing for residual base
base, is often acceptable, provided there are no strong loads, a design end-bearing capacity of around 0.4qc , inde-
stratigraphic changes within the wider range. pendent of diameter, appears reasonable. This may turn out
Chow (1997) assembled a database of high-quality pile to be conservative in cases where high residual stresses can
load tests, and her data for the end-bearing resistance of be justified, for example for jacked piles or where the
closed-ended piles driven into sand are shown in Fig. 12. transient base pressures mobilised during pile driving can be
The values of cone resistance have been obtained by aver- shown to be a high proportion of qc .
aging over 1.5d relative to the pile base, and the ultimate The vagaries of data interpretation, and the inevitable
end-bearing resistance, qbu , is that mobilised at a pile base subjectivity involved, are well illustrated in Fig. 13(b), which
displacement of 0.1d. The design curve proposed by Jardine shows a recent reinterpretation of the same database by
& Chow (1996) is indicated, and is expressed as White (2003). The reinterpretation includes:

1.0
KA (Franki) E (O,d)
0.9
KA (Cone) A (C,d)
0.8

0.7 D (C,d) G (C,d)


Design curve from
Jardine & Chow (1986)
0.6 DK (C,j) LB (C,j)
qbu/qc

0.5 HP (C,d) HT (C,d)


0.4 AK (C,j) AK (C,d)
0.3
S (C,d) BG (C,d)
0.2
Key to individual pile tests
0.1 from Chow (1997)
0
0 0.2 0.4 0.6 0.8 1.0
Pile diameter: m

Fig. 12. Normalised end-bearing capacities for closed-ended piles from Chow (1997)
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 857

1.0 estimation of qc from SPT data, and insufficient base


Driven pile (suspect data point) displacements to estimate qbu .
0.9

0.8 In the light of these caveats and adjustments, evidence for a


Residual load
corrections
significant diameter effect is unconvincing, provided appro-
0.7 priate averaging of the cone resistance is undertaken and
0.6 Suggested design value ‘ultimate’ base capacity is assessed in terms of relative
displacement (proportion of pile diameter) not absolute dis-
qbu/qc

(diameter independent)
0.5
placement. Particular care should be taken in strongly strati-
0.4 fied soils, for example where piles are driven through weak
material and penetrate only 1 or 2 diameters into a dense
0.3
Jacked piles sand layer. For such cases the design cone resistance needs
0.2 to be weighted to reflect the overlying weaker material
0.1
(Meyerhof, 1976; Meyerhof & Valsangkar, 1977).
Vibro-driven
Corresponding end-bearing data for open-ended piles are
0 shown in Fig. 14. There are many fewer data points, and
0 0.2 0.4 0.6 0.8 1.0
Pile diameter: m
they are very sparse for diameters in excess of 1 m, which is
(a) the main area of interest for offshore applications. Again
there appears to be a decreasing trend of normalised end-
1.2 bearing resistance with increasing pile diameter. However,
1.1 scrutiny of the data reveals that:
1.0
(a) the data for piles of diameter 1 m (HO: Kusakabe et
0.9 al., 1989) and 1.2 m (K: Ishihara et al., 1977) are
0.8 projected from tests where the base movement was only
0.7 0.5% of the pile diameter
qbu/qc

(b) the data point for the pile of 2 m diameter (T: Shioi et
0.6 qc reassessed (shallow
penetration of sand layer)
al., 1992) has been normalised using a cone resistance
0.5 of 35 MPa, whereas the pile tip was very close to the
0.4 top of a much softer stratum (see Fig. 15).
0.3 Possible clay
Zero residual Certainly for design a much more conservative value of qc
0.2 load observed layer at base
would be adopted in this case, possibly as low as 10 MPa.
0.1 qc estimated from SPT In order to arrive at an acceptable design approach for
0 large diameter open-ended piles, it is necessary to consider
0 0.2 0.4 0.6 0.8 1.0 the mechanics of the soil plug (Fig. 16). If the soil plug
Pile diameter: m starts to slip relative to the pile, then the shear stresses
(b)
around the plug, which are themselves a function of the
Fig. 13. Commentary on database of pile end-bearing capacity: average vertical effective stress in the plug, will lead to an
(a) annotated database from Chow (1997); (b) alternative exponential growth in the vertical stress within the soil plug.
interpretation of data by White (2003) It may be shown (Randolph et al., 1991) that the available
end-bearing resistance may be expressed as
qbplug
(a) adjustment of design cone resistance values to allow  e4 hp =d i (19)
for partial penetration into a dense sand layer [Kallo  v9base
(KA: de Beer et al., 1979); Lower Arrow Lake where hp is the height of the soil plug, di is the internal pile
(E: McCammon & Golder, 1970)] diameter, and  v9base is the ambient vertical effective stress at
(b) correction to include residual base load [Drammen (D: the base of the plug (taken as ª9hp ). As for the external
Gregersen et al., 1973)] shaft friction, the ratio  ¼ s = v9 may be expressed as
(c) reservations on the quality of the data, such as K tan , where  is the interface friction angle. Although the

0.6

DK (O,d) G (O,d)
0.5

DK (O.d) HO (O,d)
0.4

Limited base
qbu/qc

0.3 movement (0.5 % of d) CH (O,d) K (O,d)

0.2
T (O,d) CR (O,d)

0.1 Key to individual pile tests


Overestimated qc from Chow (1997)
0
0 0.5 1.0 1.5 2.0
Pile diameter: m

Fig. 14. Normalised end-bearing capacities for open-ended piles from Chow (1997)
858 RANDOLPH
TP pile wall. Near the tip, however, relative slip will occur
0 10 20 30 40 50 60 24.4m
during loading, owing to compression of the soil plug.
Lehane & Randolph (2002) have considered the separate
response of the soil plug, soil below the pile base, and pile–
soil interaction around the annular tip, in order to establish
minimum values for the end-bearing capacity of open-ended
10 piles in sand. Their recommendations are shown in Fig. 17,
based on conservative assumptions that ignore any increase
in the stress ratio, K, near the pile tip or densification of the
Depth below seabed: m

soil within the plug or beneath the pile base (together with
any associated residual stress systems). Even moderate re-
20 laxation of these assumptions suggests that a design end-
bearing capacity of qb /qc  0.2 is reasonable, and such a
value is broadly consistent with the database in Fig. 17,
taking account of limitations in the data plotted for the piles
TP of diameter greater than 1 m. Results of centrifuge model
55.0m tests in dense sand also support this as a lower bound design
30
value (Bruno, 1999; De Nicola & Randolph, 1999).
Pile tip

35 MPa assumed
but could be  10 MPa Shaft friction
40
Since the pioneering work of Vesic in the 1960s (Vesic,
1967, 1970), it has been realised that in sand and other soils
Fig. 15 Cone resistance profile for Tokyo Bay pile test (after of high permeability the magnitude of shaft friction at a
Shioi et al., 1992) given depth can reduce as the pile is driven further, with the
net effect that the average friction along the pile shaft can
reach a limit and even reduce as the pile embedment in-
creases. However, this effect has only recently been quanti-
Soil plug fied, through the carefully instrumented pile tests undertaken
Pile wall
by the research group at Imperial College. The phenomenon
of ‘friction degradation’ is illustrated in Fig. 18 (Lehane
et al., 1993), with profiles of shaft friction measured in the
σv′ three instrument clusters at different distances (h) from the
tip of a pile 6 m long and 0.1 m in diameter, as it is jacked
τ  βσv′ into the ground. For comparison, the cone profile is plotted
on the same scale, but with qc factored down by 100. The
γ′
shaft friction measured at h/d ¼ 4, in particular, follows the
shape of the qc profile closely, allowing for differences in
cone and pile diameter. Comparison of the profiles from the
instrument clusters at h/d ¼ 4 and h/d ¼ 25 shows that the
σv′  dσv′
friction measured at the latter position is generally less than
50% of that measured close to the pile tip.
The physical basis for friction degradation is the gradual
densification of soil adjacent to the pile shaft under the
cyclic shearing action of installation. This process is en-
hanced by the presence of crushed particles from the pas-
Fig. 16. Equilibrium of soil element within soil plug sage of the pile tip, which gradually migrate through the
matrix of uncrushed material (White & Bolton, 2002). The
far-field soil acts as a spring, with stiffness proportional to
value of K may be as high as unity close to the pile tip G/d (where G is the soil shear modulus), so that any
(Paik & Lee, 1993; De Nicola & Randolph, 1997; Lehane &
Gavin, 2001), it has been found to decay rapidly along the
length of the soil plug. Minimum values of  may be 0·4 Driven piles
deduced from Mohr’s circle considerations, and lie in the
wb/d  0·2
range 0.15–0.2 for typical soil friction angles (Randolph et
al., 1991). 0·3
wb/d  0·15
From equation (19), the available end-bearing pressure
rises rapidly with the plug length, so that lengths of only a wb/d  0·1
qbu/qc

few diameters can provide sufficient internal resistance to 0·2


ensure ‘plugged’ failure mode under static loading, regard-
less of the pile diameter. This contrasts with recommenda-
tions of Hight et al. (1996) and Jardine & Chow (1996), 0·1
Bored pile (Lee & Salgado, 1999)
where driven piles of diameter greater than
2(Dr  0.3) metres (with Dr the relative density, expressed
as a fraction) are assumed not to plug. 0
0 0·2 0·4 0·6 0·8 1·0
Part of this divergence of opinion revolves around the Relative density, Dr
semantics of ‘plugged’ or ‘unplugged’. Here, the term
‘plugged’ is restricted to the situation where, during static Fig. 17. Normalised end-bearing capacity for open-ended piles
loading, the top of the soil plug does not slip relative to the (after Lehane & Randolph, 2002)
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 859
Local shear stress: kPa soil annulus is proportional to G/d, and hence the reduction
0 10 20 30 40 50 60 in radial stress resulting from any contraction of soil in the
0
interface layer will be higher for smaller-diameter piles.
The MTD method, derived from the Imperial College field
studies and database of high-quality pile tests, expresses the
1
shaft friction for driven piles in sand as
"  :  : #
0 38
qc  v90 0 13 d
2 s ¼ þ ˜ r9d tan cv (21)
Depth of instrument: m

45 pa h

Cone resistance
where pa is atmospheric pressure (100 kPa) and ˜ r9d is a
3 qc /100 (relatively small) stress increase due to dilation during load-
ing (Jardine & Chow, 1996). The minimum h/d is taken
h/d  25 conservatively as 4 (in the absence of data at lower h/d
4 ratios), and for open-ended piles the diameter, d, is replaced
by the equivalent diameter, deq . The method adopts a power
law degradation, rather than an exponential decay, but this
h/d  14 leads to similar shapes of shaft friction profiles.
5
A comparison between the MTD and the present method,
h/d  4 using equation (20) with K max ¼ 0:01qc = v90 , is provided in
Fig. 19, for a 1 m diameter open-ended pile driven 40 m into
6 sand. The main difference is close to the pile tip, where the
Fig. 18. Measured profiles of shaft friction (Lehane et al., 1993)
MTD method yields identical values of shaft friction for
open- and closed-ended piles (for h/deq , 4). The present
method gives different maximum values of shaft friction,
densification close to the pile results in reduced radial effec- dictated by Kmax , and it is suggested that Kmax is increased
tive stress. The operative value of G will be high, as the soil to 0:015qc = v90 for closed-ended piles in view of the higher
is heavily overconsolidated having moved through the zone normalised end-bearing resistance. The shaft friction ratio
of high stress close to the pile tip during installation and is between open and closed-ended piles implied by the two
being unloaded. methods is quite similar, with an average ratio of around
The incremental volume change, and hence reduction in 0.7, although the MTD method gives a ratio that decreases
:
radial effective stress, is likely to depend on the current from unity at the pile tip down to (deq /d)0 38 (typically
stress level, with greater changes at higher stress levels. This 0.65) whereas the present method gives an increasing ratio
suggests an exponential variation of radial stress along the as K approaches Kmin for both geometries. The average ratio
pile shaft of the form (Randolph et al., 1994) of 0.7 may be compared with the API (1993) design
 r9 s recommendation of 0.8, but also with recent experimental
K¼ ¼ ¼ K min þ (K max  Kmin )e h=d studies that show a much lower ratio of just under 0.5 (Paik
 v90  v90 tan cv et al., 2003).
(20)
where Kmax may be taken as a proportion of the normalised
cone resistance, typically 1–2% of qc = v90 , and Kmin lies in Shaft friction in tension and compression
the range 0.2–0.4, giving a minimum friction ratio, s = v90 , The tensile capacity of piles in sand has been found to be
of 0.1–0.25 (Toolan et al., 1990). The coefficient  may be less than the shaft capacity measured in compression, and
taken in the region of 0.05 for typical pile diameters, most design guidelines include a reduction of 10–30% to
although there are some indications that the value decreases allow for this (API, 1993). Two factors were identified by
as the pile diameter increases and vice versa. Indeed, much De Nicola & Randolph (1993) that contributed to lower
higher values of  are needed to match data from centrifuge tensile shaft friction: the first was a reduction in effective
model tests (Bruno, 1999), although scaling problems related stress levels adjacent to the pile compared with loading in
to the spring stiffness of the surrounding soil may occur for compression (even for a rigid pile), and the second was the
centrifuge modelling of piles in sand (Fioravante et al., Poisson’s ratio reduction in diameter (and consequential
1999; Fioravante, 2002).
Other key variables that affect the rate of degradation
include Shaft friction: kPa
0 100 200 300 400
(a) the unloading modulus of the soil (probably close to 0
the small strain value, G0 ), with higher unloading Open-ended pile: L  40 m, d  1 m, ρ  0.1
5
modulus leading to more rapid degradation qc  5  1z MPa, γ′  11 kN/m3, δcv  28°
(b) the number of ‘shearing cycles’ per diameter of 10 Kmax  0·01qc /σ′v0, Kmin  0·3
advance (or blow count for driven piles). 15
Present method:
Assuming that cyclic stress reversal is the major trigger for
Depth: m

20 µ  0·03, 0·05 and 0·07


compressive volumetric strain in the shearing zone, the rate
25
of degradation should be very low for continuous jacking
(De Jong & Frost, 2002), and maximum for driven piles 30
with high blow count. For intermittently jacked piles such as MTD method
35
the Imperial College instrumented pile tests, the rate of
degradation would be intermediate, lower than most driven 40
piles, although this may be compensated for by scale effects 45
associated with the small diameter of the pile. As noted by
Fioravante (2002), the unloading stiffness of the surrounding Fig. 19. Example profiles of shaft friction for driven pile in sand
860 RANDOLPH
reduction in radial effective stress). These two effects were Cone resistance, qc: MPa
quantified for piles fully embedded in sand, by the expres- 0 20 40 60 80 100
sion 20
   Simplified design profile
(Qs )tens 100
 1  0:2 log10 (1  8 þ 25 2 ) (22) 25
(Qs )comp L=d
30
where Qs is the shaft capacity and ¼
p (L/d)(Gave /Ep )tan ,
with Gave , Ep and
p being respectively the average soil

Depth: m
35
shear modulus, Young’s modulus of an equivalent solid pile
and Poisson’s ratio for the pile. 40
The two factors that contribute to reduced tensile capacity
tend to compensate as the pile aspect ratio increases, with 45
the average change in effective stress level decreasing and
the effect of Poisson’s ratio contraction increasing. This is 50
shown in Fig. 20 where, for a typical modulus ratio of
Ep /Gave ¼ 400, the shaft capacity ratio is 0.8 for a range 55
of L/d. Even for quite wide extremes of Ep /Gave , the shaft
capacity ratio remains within 0.7–0.85. Fig. 21. Cone resistance profiles from Euripides site
Although other effects, such as local stress changes due to
dilation, will influence the shaft capacity ratio, the expres-
sion in equation (22) provides a reasonable design basis for The resulting profiles of pile resistance are shown in Fig.
assessing the reduced shaft capacity for loading in tension, 22, compared with the measured loads at a pile displacement
compared with that for loading in compression. of 0.1d. The test pile was initially driven to depths of
30.5 m, 38.7 m and 47.0 m, with compression and tension
tests being conducted at each level. The pile was then
Euripides pile test extracted and driven without pause to 46.7 m penetration,
A major field investigation of driven pile capacity in after repairing some of the damaged instrumentation. Over-
dense sand, the Euripides pile test, was undertaken in the all, the agreement between either prediction method and the
1990s, funded by a number of companies operating in the test data appears reasonable over the range where the
offshore area and managed by Fugro BV. The data are now instrumentation survived the driving process. However,
in the public domain and are in the process of being the measurements show very low friction mobilised over the
published: a brief comparison of measured and predicted depth range 22–30 m, and a need is indicated for some
pile resistance is presented here. The pile test comprised a refinement of the average cone resistance (possibly averaging
heavily instrumented pile, 0.76 m in diameter and 35 mm over a greater distance above pile base level). At greater
wall thickness, driven open-ended into very dense sand. depths, the MTD and exponential decay methods give shaft
Cone resistance profiles are shown in Fig. 21, and it may be friction values (or gradients of axial load) that lie respec-
seen that, in spite of some variability in the cone resistance tively slightly above and below the measurements. Predic-
below 30 m, the average qc rises to between 60 and 70 MPa. tions of end-bearing resistance appear reasonable, with close
The cone resistance over the upper 22 m is very low, and agreement between the MTD method and measured base
most of the test pile capacity was generated below that load at the intermediate depth.
level. The average shaft friction ratio between tensile and com-
The pile resistance mobilised in the dense sand has been pressive load tests ranged between 0.6 and 0.9, but with no
estimated using the MTD method, and also that presented clear pattern among the four separate load test depths.
here, adopting the design qc profile shown in Fig. 21. The Further assessment of any residual loads (or load cell zeros)
MTD method gives an end-bearing ratio of qb /qc ¼ 0.17, may lead to some revision of these estimates, but the range
whereas a ratio of 0.2 has been adopted for the alternative spans that shown in Fig. 20 based on the approach of De
method. Values of Kmax , Kmin and  have been taken as Nicola & Randolph (1993).
0:01qc = v90 , 0.3 and 0.05 respectively, whereas the value of
cv measured in ring shear tests was 308.

0·9 Pile load: compression: MN


Ep/Gave  800
0 5 10 15 20
0
400 Field data: location 1
0·8 Field data: location 2
Shaft capacity ratio

10
200 MTD method
Exponential decay: µ  0.05
0·7
20
Depth: m

0·6 30
tanδ = 0·5
νp = 0·3 40
0·5
0 10 20 30 40 50 60
Pile aspect ratio, L/d 50

Fig. 20. Ratio of shaft capacity in tension and compression (De Fig. 22. Measured and calculated load distributions for Eur-
Nicola & Randolph, 1993) ipides pile tests
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 861
Summary Poulos (1998), who compares the relative merits of static,
We now have a much clearer picture of changes that Statnamic and dynamic tests
occur in the soil around piles driven into sand, even though Further discussion here is confined to the interpretation of
our design calculations rely heavily on empirical correla- dynamic pile tests, which have the potential to be extremely
tions. Use of the cone resistance provides a better quantita- cost-effective, but where commercial practice has fallen
tive basis for these correlations, but we need to review how behind significant advances in modelling the dynamic inter-
best to average qc , taking account of the pile diameter, the action between pile and soil. This has contributed to a
displacement needed to reach steady-state conditions, and growing scepticism regarding the confidence that may be
the soil stratigraphy. The effect of pile diameter on design placed in such tests. The discussion will focus on modelling
end-bearing resistance, or on the plugging of pipe piles, is the inertial (or radiation) damping from the soil surrounding
an area of apparent divergence between science and empiri- the pile, and on treatment of open-ended piles, but a brief
cism, which needs to be resolved. Analysis of the soil plug description of the main principles is included first.
response suggests that the (static) end-bearing resistance of A schematic diagram of the main features of dynamic pile
pipe piles that do not plug during driving may be taken testing is shown in Fig. 23. The test comprises monitoring
conservatively as about half that for a comparable closed- of strain and acceleration by means of instruments attached
ended pile, but that design values of qbu /qc may be strongly to the pile shaft above ground, and preferably at least two
affected by the magnitude of residual base pressure, qb0 , or diameters below the pile head, during blows applied by
the extent to which the base response has been pre-stiffened standard piledriving hammers or large drop-weights. Nor-
during the installation process. mally pairs of instruments are located at opposite sides of
Modern design methods must take account of friction the pile in order to minimise effects of bending, and the raw
degradation, but further work is needed in order to explore data are converted to force, F (by multiplying the measured
how the rate of degradation is affected by pile diameter, strain by the cross-sectional rigidity, EA, of the pile) and
method of installation (particularly blow counts during driv- velocity, v (by integrating the accelerometer data).
ing), and soil modulus. Initially, the stress wave travels down the pile unimpeded,
The design approaches considered here are conservative in during which time the measured force and velocity are
two respects. The true ultimate base resistance will exceed directly proportional to each other: that is, F ¼ Zv, where Z
the design value based on a limited displacement, with the is the pile impedance, EA/c, with c being the wave speed in
average end-bearing resistance ultimately approaching the the pile. Any resistance to movement of the pile, for
cone resistance. Secondly, recent studies have suggested that example due to shaft friction, or change in pile section
the shaft friction of piles in sand shows significant increase (more precisely, impedance) including at the pile tip, will
with time (Chow et al., 1998), with gains of 50–100% cause an upward-travelling wave to be propagated back up
possible. Although the resulting shaft friction may prove towards the pile head. Monitoring two properties (force and
somewhat brittle, and so should not be considered in con- velocity) allows separation of downward (subscript, d) and
junction with large displacements to mobilise the base upward (subscript, u) components of the stress wave, accord-
resistance, further understanding and quantification of this ing to
phenomenon would be valuable.
Fd ¼ 0:5(F þ Z
)
Fu ¼ 0:5(F  Z
) (23)
DYNAMIC PILE TESTING It may be shown that, for sufficient magnitude of dynamic
Overview force, the total soil resistance encountered by the pile during
Load tests to verify capacity are an essential part of most the passage of the stress wave down and up the pile is equal
piling contracts, reflecting the relatively high level of uncer- to the algebraic sum of the initial downward-travelling
tainty in predictive methods. Traditional static load tests, (maximum) force plus the upward-travelling force that ar-
using kentledge or reaction systems, undoubtedly provide the rives back at the instrumentation at a time 2L/c after the
most precise method of evaluating the load–displacement initial maximum (Rausche et al., 1985; Randolph, 1990).
response, with minor limitations in terms of interaction Simplified methods have been proposed for estimating the
(mainly affecting the pile stiffness) and loading rate effects. static component of the total soil resistance mobilised during
However, static pile tests are relatively expensive, and also a dynamic test. However, the most reliable approach to
give limited information on the distribution of resistance analyse dynamic tests is by computer simulation of the
along the pile unless it is instrumented. For large diameter
cast-in situ piles, external reaction systems become prohibi- v0
tively expensive, and alternative devices such as Osterberg Time
cells (Osterberg, 1989), which use part of the pile itself to Instrumentation:
accelerometers 0 2h/c 2L/c
provide reaction, offer a more effective means of measuring strain cells
capacity. An example application of this technique is pre- h
sented later.
Dynamic pile tests, using high-energy piling hammers,
Reflections
provide an alternative to static loading tests, at a cost that is Shock Wave
from shaft
typically two orders of magnitude lower, but require sophis- wave speed,
resistance
c
ticated numerical simulation of the measured stress waves in
order to interpret the test. Alternative methods of testing, at L 1 Upward
intermediate loading rates (and costs) have appeared over Pile travelling
the last decade, in particular the Statnamic test (Berming- impedance: (reflection
Z  EA/c from base)
ham & Janes, 1989), in which a fast-burning fuel is used to Downward
accelerate a mass away from the pile, thereby loading it in travelling
Depth
reaction. Reviews of various alternative methods for pile
load tests have been given by England & Fleming (1994),
who focus on different procedures for static load tests, and Fig. 23. Schematic of stress wave travel down pile
862 RANDOLPH
pile–soil interaction, endeavouring to match computed and the range 0.2–0.5, and m is 0.3–0.5 for sand, and as high as
measured stress-wave signals. The computer model is 2 or 3 for clays.
the dynamic equivalent of the load transfer approach for A typical response of the model described by equations
static (non-linear) analysis, with the pile modelled as a one- (25) and (26) is shown in Fig. 25(a), compared with a
dimensional elastic column and static load transfer springs corresponding result for the Smith (1960) model in (b) –
replaced by an appropriate dynamic model. note that both results are from the example application
described later. Various features are evident from this com-
parison:
Dynamic pile–soil interaction models
(a) The initial response of the continuum model is
Worldwide, the majority of dynamic pile tests are ana-
extremely stiff, and dominated by the inertial dashpot
lysed using pile–soil interaction models based on Smith
term in equation (25). As noted by Randolph & Deeks
(1960), with the soil resistance expressed as
  (1992), for typical acceleration levels the dashpot term
wp will be at least a factor of 10 larger than the spring
 ¼ Min 1, (1 þ J
)s (24)
Q term during the initial response, and here slip starts to
occur after a pile displacement of 0.3 mm, when the
where wp is the local pile displacement,  and s are the dashpot accounts for 97% of the resistance.
dynamic and static shaft friction values (or equivalent in (b) By contrast, the Smith model with a standard ‘quake’
end-bearing pressure), Q is the displacement, or ‘quake’, to of 2.5 mm and damping value, J, of 0.2 s/m shows a
mobilise the limiting static resistance, and J is a damping much more gradual development of resistance: the
parameter multiplying the local pile velocity, v. immediate consequence of this is that the profile of
In the Smith model, separate viscous and inertial effects shaft friction deduced using the continuum model will
are lumped into a single parameter, and the model does not be offset to significantly greater depths, compared with
differentiate properly between conditions prior to pile–soil the Smith model (as the pile will take 1–3 ms to move
slip and those after slip. A more rational model is shown in 2.5 mm, during which time the stress wave will have
Fig. 24, as proposed by various researchers in the 1980s travelled 5–15 m).
(Simons & Randolph, 1985; Lee et al., 1988) and subse- (c) The dynamic amplification of the static shaft friction is
quently implemented in many research-oriented, but few affected by the velocity exponent, n (here taken as 0.2
commercial, codes. In what will be referred to here as the in the continuum model, and unity in the Smith model);
continuum model, the far-field soil response may be mod- a non-linear version of the Smith model is straightfor-
elled accurately using an elastic spring in parallel with a
dashpot representing inertial damping (Novak, 1977), with
the shaft response expressed as 1.5
Total
 
w

G þ < lim (25) 1.0


Static limit
d
s
Inertial dashpot
where vs is the shear wave velocity in the soil. Here, w and 0.5
Shear stress, τ/τs

Elastic spring
v are the displacement and velocity of the soil immediately
adjacent to the pile–soil interface, rather than the pile, Viscous dashpot
0
which is an important distinction.
The interface itself is modelled using a limiting shaft
friction that is velocity dependent, such as 0.5
"   n#
˜v 1.0
lim ¼ s 1 þ m (26)

0
1.5
where m and n are viscous parameters and ˜v is the relative 0 5 10 15 20 25
(or slip) velocity between pile and soil, normalised by v0 Pile displacement: mm
(taken for convenience as 1 m/s). Studies by Litkouhi & (a)
Poskitt (1980) suggest that the exponent n typically lies in
2.5

2.0 Total

1.5
Pile–soil interface Pile node Static response
Shear stress, τ/τs

1.0

0.5 Viscous
Plastic component
slider Viscous
dashpot 0
Far-field
response 0.5

1.0
Elastic Inertial
dashpot 1.5
spring
0 5 10 15 20 25
Pile displacement: mm
(b)

Fig. 25. Example responses of (a) continuum and (b) Smith


Fig. 24. Model for dynamic pile–soil interaction along shaft shaft models
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 863
ward to implement, although it is rarely adopted. Even bridge across the Swan River. Fig. 27 shows a typical
so, the continuum model tends to give a more sustained geometry of the steel pipe piles, with diameter of 0.61 m
maximum resistance, as the spring component continues and wall thickness of 12.5 mm. In most piles, an annular
to rise as the inertial component decreases, whereas the diaphragm was incorporated a few metres back from the pile
formulation of the Smith model automatically gives a tip, in order to encourage the development of a fully
decline in resistance as the pile velocity decreases. plugged condition. Without the diaphragm, it was found that
the piles could be driven to large penetrations with relatively
A similar ‘continuum’ model exists for the dynamic inter-
low energy, and although the pile would undoubtedly have
action at the base of the pile (Deeks & Randolph, 1995),
responded in a plugged manner to static loading, it was not
and again the inertial component dominates the initial
possible to verify the plugged capacity dynamically. The soil
response, although not to quite to the same extent as for the
stratigraphy for the pile groups supporting the approach
shaft (Randolph, 2000).
piers and abutments is indicated in Fig. 27: it comprised
sand fill overlying alluvial soft clays and sand lenses above
a dense sand base in which the piles were to be founded.
Open-ended piles
Matching of computed and measured stress waves can be
The driving response of open-ended piles is usually
achieved by using one of the signals (such as the force) as
evaluated by lumping internal and external friction together
input, and matching the other signal (velocity). However, a
(as external friction) and considering the end-bearing resis-
better approach is to combine the signals together in the
tance on the annular base. However, the response of the soil
manner of equation (23), using the downward-travelling
plug is very different from that of the external soil, as
wave (the externally applied force) as input, and matching
energy cannot be propagated to the far-field. A way of
the upward-travelling waves (due to soil resistance and tip
modelling the soil plug, as a column of nodal masses
reflection).
interconnected by damped springs, and its interaction with
Figure 28 shows examples of matching the upward-travel-
the pile was first proposed by Heerema & de Jong (1979)
ling force wave using three approaches:
and extended using continuum-style soil models by
Randolph (1987), as shown in Fig. 26. Recent studies by (a) explicit modelling of the soil plug using the continuum
Liyanapathirana et al. (2001) have shown that this approach soil model
works well for conditions where the incremental filling ratio, (b) internal and external shaft friction lumped together,
defined as the ratio of incremental plug movement within using the continuum soil model
the pile, to the penetration of the pile (Brucy et al., 1991) (c) internal and external shaft friction lumped together,
exceeds 50–70%. using the Smith (1960) model.
All results are for identical distribution of friction down the
inside and outside of the pile, and identical base resistance
Example application
on the pile annulus. It is clear that lumping the internal shaft
Examples of dynamic testing of steel pipe piles are
friction with the external friction has had a significant effect
presented here, drawn from a recent project in Perth, Wes-
on the shape of the computed stress wave, even using an
tern Australia, involving duplication of the main freeway
identical continuum model. Also, the response computed
using the Smith (1960) model is very different from that
using the continuum model. In fact, for this particular open-
ended pile test, no good match could be obtained without
External explicit modelling of the soil plug, even by varying the
soil distribution of soil resistance.
Soil plug
The best (although not particularly close) fit to the

0.61 m
Shaft

Steel
pipe
pile
Sand
fill
Soil response
model

33 m
Pile wall

Base Soft
clay

Annular steel
diaphragm

Base response 6m
model Dense
Soil
sand
plug

Fig. 26. Modelling dynamic response of soil plug Fig. 27. Partially plugged pipe pile
864 RANDOLPH
3.0

Computed
2.5 Smith model
closed-ended
Computed
continuum model
2.0
closed-ended

1.5

Force: MN
1.0

0.5 Computed Measured


open-ended

0
0 10 20 30 40

0.5

1.0
Time: ms

Fig. 28. Matching of measured and computed upward-travelling waves

measured upward-travelling wave in Fig. 28 was obtained 8


with the capacities summarised in Table 2 (Pile 1). The
quality of fit was relatively sensitive to the internal shaft 7
friction, and reasonable fits were obtained for mobilised
6
resistances in the range 6.5–7.5 MN. The static response of Range of reasonable
fits to dynamic test
the pile was evaluated by lumping the internal shaft capacity 5 Measured
Force: MN

into the base resistance, and assuming that the resulting base
resistance of 4.91 MN would be mobilised at a displacement 4
of 10% of the pile diameter (0.06 m). The distribution of Computed
3
external shaft friction (and shear modulus) deduced from the
dynamic test was incorporated unaltered into the static 2
analysis. The resulting pile response is compared with the
results of a static load test on the pile in Fig. 29 and shows 1
surprisingly good agreement considering the rather arbitrary
0
manner of incorporating the dynamic soil plug resistance in 0 20 40 60 80 100 120
with the base response. Displacement: mm
One further example of predicting the static pile response
from a dynamic test is shown in Fig. 30, for a tension test Fig. 29. Comparison of measured and computed pile response
carried out as part of the same project. In this case the
open-ended pile was driven (with no internal diaphragm) Summary
through the upper sand fill to tip into the underlying clay, Dynamic pile tests are arguably the most cost-effective of
and a plug length of 10 m (compared with total pile penetra- all pile-testing methods, although they rely on relatively
tion of 16.5 m) was measured. The stress-wave matching has sophisticated numerical modelling for back-analysis. Theor-
been discussed by Randolph (2000), and the resulting capa- etical advances in modelling the dynamic pile–soil inter-
cities are summarised in Table 2. action have been available since the mid-1980s, but have
The computed response under (static) tensile loading, been slow to be implemented by commercial codes, most of
taking just the external shaft resistance, is compared with which still use the empirical parameters of the Smith (1960)
the measured response in Fig. 30. The measured capacity in model. In order to allow an appropriate level of confidence
tension is about 15% lower than the external shaft capacity in the interpretation of dynamic pile tests, and estimation of
deduced from the dynamic test, which is consistent with the the static response, it is high time that appropriate scientific
ratio of tensile and compressive shaft capacities from Fig. models were used for pile–soil interaction, including explicit
20. However, it should also be pointed out that the level of modelling of the soil plug for open-ended piles.
confidence in the back-analysed external shaft capacity is There will still be a need for some empiricism, mainly in
poor, with a possible range of 1–1.5 MN (mainly by redis- adjusting the static shaft friction to allow for the very high
tributing internal friction to external friction). shearing rates during a dynamic test. Modelling viscous and

Table 2. Summary of mobilised soil resistances from dynamic tests


External shaft Internal shaft Annular base Total pile
resistance: resistance: resistance: resistance:
MN MN MN MN
Pile 1 (compression) 2.43 3.97 0.94 7.34
Pile 2 (tension) 1.02 0.98 0.06 2.06
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 865

1.6
Range of reasonable
1.4 fit for external shaft
friction from
1.2 dynamic test Steel
pipe
1.0 pile
Computed
Force: MN

0.8 Measured
16·5 m
0.6 10 m

0.4
0·61 m Soil
plug
0.2
(a) (b)
0
0 10 20 30 40 50 Fig. 31 Pile group capacity and stiffness: (a) capacity dependent
Displacement: mm on conditions at pile–soil interface; (b) stiffness determined
Fig. 30. Comparison of measured and computed pile response primarily by far-field conditions
in tension
pile groups, and the trend towards piled rafts where raft
other effects needs to be placed on a more rational basis, as foundations are supported by relatively low numbers of piles,
approaches such as equation (26) are far from ideal. Finally, placed strategically in order to minimise differential settle-
it should be realised that the deduced soil resistance para- ments and bending moments in the raft. However, a case
meters will not be unique. It is therefore important when study involving foundations subjected to general vertical,
back-analysing dynamic pile tests to adopt appropriately horizontal and moment loading will also be presented,
conservative values (or distributions) of soil resistance that illustrating the importance of considering the overall founda-
still give a reasonable fit to the stress wave data, rather than tion response, rather than ‘safety factors’ on individual piles.
necessarily adopting the ‘best fit’ set of parameters. Given
the empirical nature of the soil models, the concept of ‘best
fit’ is of limited value, and a philosophy of ‘conservative,
while still consistent’ is a more prudent approach. Analysis of vertically loaded pile groups
Many different approaches may be used to analyse the
response of single piles and pile groups subjected to vertical
DESIGN OF PILE GROUPS loading. The most widespread, and most rigorous for homo-
Design philosophy geneous soil conditions, is the boundary element method
The main objective of pile group design is to ensure that (Poulos, 1968; Banerjee & Butterfield, 1981; Basile, 1999),
the foundation does not undergo excessive displacements which has the advantage that only the pile–soil interfaces
during its design life. This principle is represented by the need to discretised, rather than the full continuum. A much
serviceability limit state in design codes, but in addition simpler approach, leading to analytical solutions, is to adopt
there is normally a requirement to ensure that there is an a Winkler approximation, where values of shear stress down
adequate material factor against ‘collapse’, or ultimate limit the pile shaft, and end-bearing pressure at the base, are
state. Where piles exhibit strain-hardening, the ultimate limit taken as proportional to the local pile displacement. This
state for an individual pile is customarily (but somewhat approach has been shown to lead to good agreement with
arbitrarily) taken as the load to cause either an absolute results from more rigorous numerical analysis (Randolph &
displacement (such as 50 or 100 mm), or a relative displace- Wroth, 1978), and recent analytical studies have evaluated
ment (such as 10% or 15% of the pile diameter). the relationship between Winkler spring constant and elastic
In practice, however, for most applications the real ulti- shear modulus in closed form (Mylonakis, 2001).
mate limit state should be determined by considering the The original solution of Randolph & Wroth (1978)
interaction of the foundation with the superstructure, and adopted the idealised assumption of shear stresses around
must consider the complete foundation system (pile group, the pile that decayed inversely with radius, leading to a
with or without ground-contacting cap) rather than indivi- logarithmic decay in vertical displacements, together with a
dual piles. As for serviceability limit states, the ultimate maximum radius of influence, rm. The shear stress at the
limit state then also reverts to a displacement criterion, but pile wall is then expressed in terms of the local displace-
one that is determined largely by structural considerations. ment, w, as
The advantage of moving more towards displacement 2Gw
criteria, reducing the emphasis on the capacity of individual ¼ (27)
d
piles, is illustrated in Fig. 31: axial pile capacity depends
critically on effective stress and fabric conditions at the where
 
pile–soil interface, which are difficult to determine accu- 2rm
rately, whereas the deformation response is influenced pri- ¼ ln
d
marily by soil conditions away from the pile, which are only
slightly affected by the installation process. Simple, but The Winkler spring constant, k, relating the force per unit
robust, analytical approaches for estimating the vertical length transferred from pile to soil to the local displacement,
stiffness of piles and pile groups are readily available, as is then given by
outlined below. Hence, provided the in situ stiffness of the d 2G
soil strata can be determined, the pile group load–displace- k¼ ¼ ¼ G (28)
ment response up to working load can be estimated with w
greater accuracy than the ultimate capacity. The value of  is typically around 1.5 for floating piles, but
The discussion here will be restricted to vertically loaded increases for ‘end-bearing’ piles (with wb ¼ 0) where it
866 RANDOLPH
varies with aspect ratio, L/d, and stiffness ratio, Ep /G, Normalised displacement, w/w1o
according to (Mylonakis, 2001) 0 0·2 0·4 0·6 0·8 1·0
 0:025 "  0:6 # 0
k 2(1 þ
)G L α  0.376 α′  0.578
 ¼  1:3 1þ7 (29)
G Ep d
0.2
The resulting pile head stiffness, K ¼ P/wt, where P and wt Pile 2 Pile 2 Pile 1
are the pile-head load and ground-level displacement respec- adjusted logarithmic
tively, may then be expressed as 0.4 decay

Depth, z/L
 þ tanh(ºL)
K ¼ Ep Ap º (30)
1 þ  tanh(ºL) 0.6
The parameters  and ºL represent non-dimensional base
stiffness and slenderness ratio for the pile, expressed as L/d  20
sffiffiffiffiffiffiffiffiffiffiffi 0.8 Ep/G  500
Pb k ν  0.3
¼ and ºL ¼ L (31) s/d  3
w b Ep Ap º Ep Ap
1.0
where Pb and wb are respectively the load and displacement
at the pile base, and Ep Ap is the cross-sectional rigidity of Fig. 33. Displacement profiles down loaded and adjacent piles
the pile.
2ºL þ sinh(2ºL)
The above expression may be used to calculate the
þ 2 [sinh(2ºL)  2ºL] þ 2[cosh(2ºL)  1]
stiffness of single piles in homogeneous soil, or in multi- ¼ (33)
layered soil (essentially using equation (30) and the asso- 2 sinh(2ºL) þ 22 sinh(2ºL) þ 4 cosh(2ºL)
ciated value of  as a transfer function; Mylonakis & This elegant capturing of the reinforcing effect of each pile
Gazetas, 1998). For soils where the stiffness varies with in a group is a seminal advance in quantifying interaction
depth, alternative solutions may be developed, involving between piles. The form of equation (33) is illustrated in
Bessel functions of fractional order (Guo & Randolph, Fig. 34, where it may be seen that  converges to 0.5 for
1997). Such solutions are readily evaluated using modern long (or compressible) piles, is less than 0.5 for end-bearing
mathematical packages. piles (high ), and is between 0.5 and 1 for most floating
For pile groups, the stiffness of each pile is reduced piles (low ).
because of interaction effects, as indicated in Fig. 32. An equivalent expression to equation (33) may be derived
Mylonakis & Gazetas (1998) demonstrated that the inter- for piles of the same length, but different diameters d1 and
action factor, Æ (as defined by Poulos, 1968), must reflect d2 (and corresponding values of ,  and ºL), leading to an
not only the (assumed) logarithmic decay in displacements, interaction factor, Æ21 (proportional settlement of Pile 2
but also the reinforcing effect of the neighbouring pile. This compared with Pile 1), given by
leads to a reduction in the interaction factor, as indicated in 2
Fig. 33. For piles of the same length and diameter, the 2 º2
sh1 þ ch1
interaction factor for a given spacing, s, may then be ex- ł21 º1 º2 4º2 º1
Æ21 ¼ 2 2 º

pressed as the product of two terms representing the loga- º2  º1 1 sh2 þ 2 ch2
rithmic decay and a ‘diffraction factor’,  (Mylonakis &  3
2 º2
Gazetas, 1998), giving (1 þ tanh º1 L) 2 sh2 þ ch2  sh1  ch1 7
  º1 7
ln(rm =s)  5
Ƽ  (32) (1 þ 1 tanh º1 L)(sh2 þ 2 ch2 )
ln(2rm =d)
where the diffraction factor, , is a function of  and ºL, (34)
according to

P
1.0
Ω0
Ω  0.05
0.8
w1t αw1t
0.1
Diffraction factor, ξ

0.6 Ω  0.2
L

1 2 0.4
Ω1
d
0.2 End-bearing pile

S
0
0 0.5 1.0 1.5 2.0 2.5
Dimensionless pile length, λL

Fig. 32. Interaction between two piles for axial loading Fig. 34. Diffraction factor, î (Mylonakis & Gazetas, 1998)
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 867
where average settlement of a pile group to the dimension, B, of
ln(rm =s) the foundation and the factor of safety against ultimate
ł21 ¼ ; sh i ¼ sinh º i L and ch i ¼ cosh º i L (35) capacity. This has been explored by Cooke (1986) and more
1 recently by Mandolini (2003), who suggest that for fine-
It may be shown that the reciprocal theorem is satisfied by grained soils the average settlement is typically around 0.3–
the above expression, as Æ21 /K1 is identical to Æ12 /K2 , and 0.6% of the foundation width, B, for a factor of safety of 3.
that the above expression reduces to that from Mylonakis &
Gazetas (1998) as the pile diameters converge.
The effect of non-homogeneous soil conditions may be Piled rafts
incorporated in an approximate manner, following Randolph As may be seen from Fig. 35, for pile groups where
& Wroth (1978), whereby the load transfer parameter, , is B/L , 1 the pile group stiffness is significantly greater than
reduced by ln(r) (where r ¼ Gaverage /GL , the ratio of the stiffness of a raft foundation. Therefore, even if the pile
average shear modulus to the value at pile tip level, ranging cap rests directly on competent ground, it will contribute
between 0.5 and 1), and also the tanh(ºL) term in the little to the response of the overall foundation. Viggiani
numerator of equation (30) is multiplied by r. (2001) has referred to such foundations as ‘small’ pile
The above approach has been used to evaluate the stiff- groups, where piles are needed to ensure adequate bearing
ness of square groups of piles, from 2 3 2 up to 30 3 30, capacity, and the pile cap (or raft) can easily be made
for L/d ¼ 25, Ep /GL ¼ 1000, r ¼ 0.75 and
¼ 0.3. The sufficiently stiff to eliminate differential settlements. By
results are shown in Fig. 35(a), where the pile group contrast, for ‘large’ pile groups, with B/L . 1, the pile cap
stiffness, Kg (ratio of total applied load to average settle- will often provide sufficient margin against bearing failure,
ment), has been normalised by GL B, where B is the width of and will contribute significantly in terms of transferring load
the pile group. Plotting the normalised stiffness against the directly to the ground. The design of such foundations
normalised width, B/L, leads to an envelope of curves that hinges more on limiting the average and differential settle-
tends to the stiffness of a surface raft as B/L becomes large. ments to a acceptable level. As, for large rafts, the flexural
The stiffness envelope may also be matched closely by using stiffness will be low, the location and length of any pile
an equivalent pier approximation of the pile group (Poulos support should be chosen in order to minimise differential
& Davis, 1980; Randolph, 1994), demonstrating the robust- settlements. Design strategies for ‘small’ and ‘large’ pile
ness of calculations of pile group stiffness even with quite groups have been discussed by Viggiani (2001) and
approximate models (Fig. 35(b)). Mandolini (2003).
From a practical viewpoint, it is also useful to link the The concept of using a limited number of piles, poten-
tially loaded to near their ultimate capacity, beneath a raft
foundation in order to reduce differential settlements was
20
L/d  25
mooted by Burland et al. (1977), and also explored further
18 by Randolph (1994) and Horikoshi & Randolph (1998). The
Ep/GL  1000
16 ρ  0·75 principle behind the design approach is illustrated in Fig. 36.
s/d  2
Group stiffness, Kg/GLB

14 ν  0.3 For large rafts (B/L . 1) the stiffness of a pile group


occupying the full area of the raft will be quite similar to
12
that of the raft alone, and a more effective approach to
10 s/d 3 reducing differential settlements is to place a few piles over
8 the central region of the raft. Where a layer of soft soil
6 s/d  5 exists beneath the raft, then it may be necessary to install
short piles extending through that layer over the full raft
4 s/d  10 area, but then to use longer piles in the central 25–40% of
2 the raft area.
Raft stiffness
0 Piled rafts of this nature require careful analysis, but there
0 1 10 are a variety of analytical approaches of varying complexity
Normalised width of pile group, B/L that are now available (Franke et al., 1994; Poulos, 1994,
(a) 2001; Clancy & Randolph, 1996; Russo, 1998; Katzenbach
et al., 2000). Studies on optimising pile geometry have
20 mainly been restricted to uniform loading of the raft, and
18
16
Group stiffness, Kg/GLB

14 Stiffness of
incompressible pier
12
Equivalent pier
10
(same area and
8 length as pile group)
6
4 80% of stiffness of
2 incompressible pier

0
0·1 1 10
Normalised width of pile group, B/L
(b)
Raft foundation Piled foundation Piled raft foundation
Fig. 35. Comparison of pile group and equivalent pier stiff-
nesses: (a) normalised stiffness of pile groups; (b) comparison of Fig. 36 Transition from raft to pile group to piled raft
equivalent pier stiffness foundation
868 RANDOLPH
these have suggested that piles of length greater than about Position across centreline of raft, x/B
70% of the width of the raft are required, situated over the 0.5 0.3 0.1 0.1 0.3 0.5
central 25–40% of the raft area (Horikoshi & Randolph, 0
Krs  0.06
1998; Prakoso & Kulhawy, 2001; Viggiani, 2001). 0.05
In practice, buildings often concentrate a significant pro-

Normalised settlement, wGB/Ptotal


portion of the total load towards the outer edges of the raft, 0.10 Piled raft
core-edge loading
giving rise to concern over potential hogging of the raft. In 0.15 Piled raft
order to address that concern, a recent study by Reul & uniform loading
Randolph (2004) has also considered the case where half of 0.20
the applied load is distributed uniformly over the central 0.25
25% of the raft, and the other half is transmitted around the
raft edge. Fig. 37 shows the raft and pile geometry analysed, 0.30
together with the two alternative load distributions. 0.35
Horikoshi & Randolph (1997) proposed a raft–soil stiffness
ratio defined as 0.40
Unpiled raft Unpiled raft
rffiffiffiffiffi  0.45 uniform loading core-edge loading
E r 1 
Br tr 3
K rs ¼ 5:57 (36)
2G 1 
2r Lr Lr 0.50
(a)
where the subscript r denotes the raft properties of Young’s
modulus, E, Poisson’s ratio,
, breadth, B, length, L, and
thickness, t. The two different raft thicknesses shown in Fig. 0.025
37, of 1 m and 3 m, lead to raft–soil stiffness ratios of 0.06 Krs  1.5
Unpiled raft
and 1.5, representing practical limits of flexible and stiff 0.020 core-edge loading
rafts.
The resulting settlement patterns for piled and unpiled Normalised settlement, wGB/Ptotal 0.015
Unpiled raft
rafts are presented in Fig. 38(a) for the flexible raft (where uniform loading
the bending moments are negligible), whereas Fig. 38(b)
0.010
shows bending moment profiles for the stiff raft (where the
differential settlements are negligible, less than 1% of the
0.005
average settlement). It may be seen from this figure that Piled raft
the central pile support reduces differential settlements for core-edge loading
the flexible raft case by a factor of between 2.7 (uniform 0
loading) and 3.5 (core-edge loading), with corresponding
ratios of differential settlement to average settlement for the 0.005 Piled raft
piled rafts of 0.15 and 0.24. For the stiff raft, maximum uniform loading
(absolute) bending moments are reduced by a factor of about 0.010
2.4, at the expense of introducing greater hogging moments. 0.5 0.3 0.1 0.1 0.3 0.5
Thus, even for situations where a significant fraction of the Position across centreline of raft, x/B
applied load is concentrated towards the edges of the (b)
foundation, central pile support appears a beneficial design
approach. Fig. 38. Settlement and moment profiles across raft centreline:
(a) settlement profiles; (b) profiles of bending moment per unit
For any given design, the precise layout and geometry of length
the pile support would need to be fine-tuned by numerical
analysis. However, the principle is to optimise the design by
locating the pile support in such a way as to minimise cumulative length (number times average length) of piles
differential settlements and bending moments. In this way, used, but with the piles concentrated in the central part of
the foundation costs may be minimised without compromis- the raft (Horikoshi & Randolph, 1998; Prakoso & Kulhawy,
ing performance. Although few, if any, piled rafts have been 2001; Viggiani, 2001; Mandolini, 2003).
optimised in this way, several studies have been published
showing how conventionally designed piled foundations
would have performed adequately, and with smaller differen-
tial settlements, with as little as 30–50% of the actual General loading: a case study
The final topic covered in this paper concerns the founda-
tions for a major bridge in Vietnam, the construction of
which was funded jointly by the Vietnamese and Australian
Uniform governments. The My Thuan bridge crosses one of the two
branches of the Mekong delta, and is a cable-stayed bridge
Core:edge (50:50)
with a central span of 350 m and clearance of 37.5 m (Fig.
38 m
39). Maunsell Australia Pty Ltd were the consulting engi-
6m neers for the project, and the author was a member of the
1 or 3 m technical advisory group for the Australian Agency for
1m
International Development (AusAID). The case study is
30 m instructive in two ways, first for its use of sophisticated pile
tests in order to finalise the design of the main pile support
38 m for the bridge towers (Chandler, 1998), and second as an
illustration of the effects of redistributing load away from
heavily loaded piles, effectively allowing the ultimate limit
Fig. 37. Example piled raft showing two alternative load state to be considered in terms of allowable deformations.
distributions The main towers of the bridge are supported on founda-
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 869
stratum of dense sand deemed necessary for the tower piles.
Two pile load tests were conducted on the south bank pier,
each with a pair of Osterberg cells, as indicated. Further
load tests were conducted on the south and north tower
foundations, using a single Osterberg cell.
The use of Osterberg cells at two levels within a test pile
allows measurement of base capacity, shaft capacity over the
pile section between the two cells, and shaft capacity of the
upper pile section, by means of three test phases conducted
in sequence (Osterberg, 1989). The results of the two south
bank load tests are presented in Fig. 42. The measured base
response was consistent in both tests and showed no evi-
dence of any accumulation of soft sediments from the
construction process. The base response has been modelled
using a hyperbolic response with an ultimate end-bearing
pressure of 4.5 MPa, and an initial stiffness that corresponds
to a shear modulus of about 450 MPa.
Fig. 39. My Thuan bridge in Vietnam The values of shaft friction measured in the first load test
were below expectations, and before the second pile was
tions comprising 16 bored piles, arranged as two groups of constructed, modifications were made to the construction
eight beneath each leg of the tower, and linked by a 6 m procedures. These included steps to reduce the delay be-
thick pile cap as shown in Fig. 40. The piles are 2.4 m tween excavation and concreting, a reduction in the head of
(nominal) diameter, around 95 m long, and were constructed bentonite above river level, and mechanical scarifying of the
under bentonite; base-grouting to 5 MPa was also performed.
Under normal conditions the water depth is around 23 m, 0m
but a deep scour hole exists upstream of the bridge where
0m
two branches of the river converge. The ultimate design Water
conditions allow for migration of the scour hole, resulting in
scouring to a depth of 47 m at the tower locations. The most
critical design loads correspond to ship impact on one of the Silty clay 23 m
tower foundations, either parallel to the river (08) or at 458 (su ~ 200 kPa)
to the river, with design loads as given in Fig. 40. The
design vertical load of 315 MN corresponds to an average
40 m
load of just under 20 MN per pile. Clayey sand 42 m
A simplified soil stratigraphy is shown in Fig. 41 for the (Φ′ ≈ 38°)
south tower and also for the first piers on the southern bank. 51 m
Only limited strength data were obtained from the site Silty clay
56 m
investigation, and instead the final pile length was based on (su ≈ 300 kPa)
the results of two load tests on piles for the south bank pier,
which were deliberately extended to reach the founding 68 m

Osterberg cells 75 m
83 m
~ 60 m Dense sand
(Φ′ ≈ 40°)
5.5 m
93 m
South bank test
2.4 m piles (86·4 m)

x South pier pile (96 m)

Fig. 41. Soil stratigraphy and location of Osterberg cells

100
Middle section
(between cells)
V  315 MN
Average shaft friction: kPa

Test 2
M ≈ 320 MNm 75
H ≈ 20 MN Test 1 Shaft

50 Upper section
(above top cell)

25

95 m 20 40 60 80 100
0
Displacement: mm
pressure: MPa

Test 1
End-bearing

2·5 Simulation (RATZ)


Base
Test 2
5

Fig. 40. Pile layout for tower foundations of My Thuan bridge Fig. 42. Response measured from load tests on south bank piles
870 RANDOLPH
borehole shaft prior to casting the pile. The last step was 40
achieved by welding short strands of wire to the callipering Load case 1:
30 ship impact at 0°
tool (Fig. 43), which was then raised and lowered within the
borehole, rotating it as required in order to cover the full
20
circumference of the shaft.
As may be seen from Fig. 42, these measures led to an Elastic

Axial load: MN
10
improvement of about 30% in the shaft friction measured in analysis
Upload Download
the second load test, with design values of 55 kPa and 0
90 kPa adopted for the upper and middle zones respectively. Non-linear analysis
Making allowance for slight differences in stratigraphy be- 10 axial limit: 22 MN
tween the south bank and tower locations, and also for the
design scour depth of 47 m, the ultimate capacity of the 20
95 m long tower piles was determined as 42.1 MN. Adopting
a material factor of 0.72 (Australian Standards, 1995), and 30 Load case 2:
allowing for the buoyant pile weight of 5.4 MN, a design ship impact at 45°
40
geotechnical ultimate capacity of 24.9 MN was arrived at for 30 20 10 0 10 20 30
the 95 m long piles. Subsequent load tests on the tower Distance from pile group centroid: m
piles, using a single Osterberg cell, resulted in a measured
ultimate capacity of the upper section of pile of 26–27 MN, Fig. 44. Computed axial loads for critical load cases
and confirmed an ultimate capacity well in excess of
30 MN. at the pile group centroid for the non-linear analyses. The
Analysis of the pile group under the design load condi- lower two curves are conventional load–displacement re-
tions, using the software PIGLET (Randolph, 2003), leads to sponse for the two load cases, as the applied loads are
a distribution of loads among the piles as shown in Fig. 44. factored up proportionally to their full values and with an
Assuming elastic response of the piles, it is found that three imposed axial load limit of 22 MN on individual piles. Final
piles under load case 1, and six piles under load case 2, displacements are 280 mm (load case 1) and 250 mm (load
exceed the design capacity of 24.9 MN. However, non-linear case 2), compared with the elastic value of 74 mm for both
analysis, with the axial load limited to a notional design cases.
capacity, allows the loads to be redistributed among the piles From a design perspective, more useful information is
at a cost of increased deformations. The load distributions given by the upper two curves in Fig. 45, which give the
for the case where the axial loads have been limited to final vertical displacements for different magnitudes of the
22 MN are shown in Fig. 44. For each load case, only four limiting axial load. This shows the more critical nature of
piles remain with axial loads less than the imposed limit of load case 2, where displacements start to increase signifi-
22 MN, suggesting that the pile group is close to failure in cantly once the axial load limit is reduced below 28 MN.
the sense of no longer being able to find a load distribution For the actual design limit of 24.9 MN the vertical displace-
in equilibrium with the applied loading. ment is 140 mm for this load case (and 76 mm for load case
Figure 45 shows the development of vertical displacement 1: little more than the elastic value).
An appropriate design strategy for bridge foundations
involves an iterative interaction with the structural design
engineers, in order to determine what displacement of the
foundation constitutes an ultimate limit state. In the present
case this allows confirmation that the geotechnical design
limit of 24.9 MN (and the resulting displacement of
140 mm) is indeed adequate, and essentially weights the
definition of ultimate limit state more towards a deformation
criterion, rather than the geotechnical capacity of individual
piles in the group.

35
Axial capacity:

Final displacement as
Load
30 function of axial
case 2
MN

capacity
25
Load
20 case 1

15 Load
Average axial

case 2 Load-displacement
load: MN

10 response with axial


Elastic capacity of 22 MN
5

0
0 0·05 0·10 0·15 0·20 0·25 0·30
Vertical displacement of pile group centriod: m

Fig. 43. Tool to scarify borehole shaft prior to concreting Fig. 45. Non-linear response of pile group under design loads
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 871
Summary deserve closer scrutiny, particularly where design requires
In contrast to other areas of pile design, the science of extrapolation outside the existing database.
predicting the deformation response of pile groups is well Overall, significant uncertainty exists in our ability to
advanced, with a variety of methods available, differing in estimate the axial capacity of individual piles, and hence we
complexity and degree of rigour. Under vertical loading, need to adapt our design and contracting strategies to en-
quantification of the reinforcing effect of piles in moderating courage the testing of piles in order to fine-tune designs,
interaction factors is a major improvement, and the approach allowing more optimistic material factors as more precise
leads to estimates of pile group stiffness that are consistent estimates of pile capacity become possible. Dynamic pile
with simple analogues based on equivalent pier or raft testing has an important role in this respect, but inertia
approximations. within the industry has restricted the adoption of more
For piled rafts, the benefits of central pile support for scientific models of pile–soil interaction.
(primarily) raft foundations have been documented widely We should capitalise on the ductile response of most piled
both for uniformly loaded pile groups and for the more foundation systems, weighting design criteria more towards
demanding case where a significant proportion of the loading allowable deformations, with reduced dependence on the
is concentrated at the raft edges. As illustrated here, even in capacity of individual piles. Efficient design of piled rafts
that case, substantial reduction in differential settlements (for may allow the piles to operate close to their geotechnical
flexible rafts) and absolute magnitude of bending moments design capacity under working conditions, with their main
(in stiffer rafts) is possible using central pile support. purpose being to minimise differential settlements and bend-
The trend towards design based on allowable deformations ing moments in what is primarily a raft foundation.
may also be appropriate for ‘small’ pile groups, such as for Despite adverse comments by some of the pioneers of soil
supporting bridge piers, which are subjected to combined mechanics, there is a significant role for scientific methods
vertical and horizontal loading. Non-linear analysis allows in pile design. If we are to continue to attract the best of
consideration of the consequences of limiting axial (or each new generation of engineers, then we must incorporate
lateral) loads on individual piles to a notional design value, such science in our teaching and our practice, using empiri-
and hence, in conjunction with structural considerations, cal approaches to validate and calibrate, but not replace,
permits even the ultimate limit state to be weighted more scientific theory.
towards allowable deformations.
The key soil parameter required for predicting the defor-
mation response of pile groups is the shear modulus, and ACKNOWLEDGEMENTS
back-analysis of the performance of large pile groups sug- I would like to gratefully acknowledge support and con-
gests that the relevant value is close to the small-strain tributions from a wide range of people during preparation of
modulus, G0 (Mandolini & Viggiani, 1997). This may be this paper: first and foremost my wife, Cherry, and sons,
measured relatively easily using modern methods such as Nick and Tom; mentors at key stages of my career, includ-
seismic cone tests or the small strain response of high- ing John Burland, Peter Wroth, Andrew Schofield and John
quality samples. For piled rafts, with fewer piles and higher Booker; colleagues Martin Fahey and Barry Lehane from
average shear stresses in the soil, some factoring of Go will UWA, and Carl Erbrich from Advanced Geomechanics;
be necessary, and further analysis and field evidence is collaborators who have supplied data and taken time to
needed in order to quantify the reduction. The use of a correspond on some of the issues, particularly David White
modified hyperbolic response to bridge between the initial but also Antonio Alvez, Fiona Chow, George Mylonakis and
stiffness (derived using G0 ) and the ultimate capacity ap- Oliver Reul; and finally past and present staff and students
pears a promising approach (Mayne & Poulos, 2001; Mayne, in the Geomechanics Group at UWA who have made it such
2003). Pile load tests carried out near the start of the piling a fun place to work. I am also grateful to AusAID for
contract provide a useful means of verifying, or fine-tuning, permission to use data from the My Thuan bridge project.
the design, not just in terms of pile capacity, but also in
respect of shear modulus values deduced from the load–
displacement response at different load levels. NOTATION
A area (Ap for cross-sectional area of pile)
Br width of raft foundation
CONCLUDING REMARKS c wave speed in pile
Einstein once commented along the lines that: ch coefficient of consolidation for horizontal flow
Dr relative density
‘‘Experimental data are believed by everyone, except the d diameter of pile (subscripts eq for equivalent, i for
person who did the experiment; while theory is believed by internal)
nobody, except the person who developed it.’’ E Young’s modulus (subscripts b for soil at pile base, p for
pile, r for raft)
This comment is particular pertinent to pile design, where F dynamic axial force in pile
we need to be vigilant over the quality of data relied upon G shear modulus of soil
in empirical approaches and to continue to improve our h distance measured upwards from pile tip (subscript p
analytical methods. With this in mind, the aim of this paper for soil plug)
has been to explore advances in scientific approaches to pile Ir rigidity index, G/su
design, and the extent to which we still need to rely on J Smith damping factor
empirical correlations. K compressibility factor, horizontal stress ratio (subscript
The scope has been deliberately restricted, partly because 0 for in situ value), pile head stiffness
of space limitations, but also reflecting the author’s interests Krs non-dimensional raft-soil stiffness ratio
k load transfer factor, permeability (with subscript h for
and experience. Summary comments have been provided at
horizontal flow)
the end of each topic and will not be repeated here. L pile length, (subscript r for raft length)
In estimating the axial capacity of piles, areas have been m non-dimensional rate factor
identified where ‘empirical’ trends from measured perform- n exponent
ance do not appear to be supported by ‘science’ in the form P load at pile head
of conceptual or analytical models. These areas therefore PI plasticity index
872 RANDOLPH
p mean stress (subscripts i for initial, o for in situ, a for Bermingham, P. & Janes, M. (1989). An innovative approach to
atmospheric pressure) load testing of high capacity piles. Proceedings of the interna-
Q Smith quake, capacity (subscript s for shaft) tional conference on piling and deep foundations, London, pp.
q pressure (subscripts b for pile base, c for cone 409–413.
resistance) Bond, A. J. & Jardine, R. J. (1991). Effects of installing displace-
R yield stress ratio ment piles in a high OCR clay. Géotechnique, 41, No. 3, 341–
Rf pile capacity reduction factor 363.
r radial coordinate (subscript m for maximum radius of Briaud, J.-L., Tucker, L. M. & Eng, E. (1989). Axially loaded 5 pile
influence) group and single pile in sand. Proc. 12th Int. Conf. Soil Mech.
req equivalent pile radius Found. Engng, Rio de Janeiro 2, 1121–1124.
St soil sensitivity Brucy, F., Meunier, J. & Nauroy, J.-F. (1991). Behaviour of pile plug
su undrained shear strength in sandy soils during and after driving. Proc. 23rd Annual
T, T  non-dimensional time factors (subscript eq for Offshore Technology Conf., Houston, 145–154.
equivalent) Bruno, D. (1999). Dynamic and static load testing of driven piles in
t time sand. PhD thesis, The University of Western Australia.
tr thickness of raft Burland, J. B., Broms, B. B. & De Mello, V. F. B. (1977). Behav-
u pore pressure (subscripts o for in situ, max for iour of foundations and structures. Proc. 9th Int. Conf. Soil
maximum value) Mech. Found. Engng, Tokyo 2, 495–546.
v velocity (subscript s for shear wave velocity in soil) Bustamante, M. & Gianeselli, L. (1982). Pile bearing capacity by
w settlement (subscript t for pile head) means of static penetrometer CPT. Proc. 2nd Eur. Symp.
z depth co-ordinate Penetration Testing, Amsterdam, 493–499.
Æ shaft friction correlation factor, interaction factor Cao, J., Phillips, R., Popescu, R., Audibert, J. & Al-Khafaji, Z.
between piles (2002). Excess pore pressures induced by installation of suction
 shaft friction correlation factor caissons in NC clays. Proceedings of the international confer-
ª unit weight (9 for effective, subscript w for water) ence on offshore site investigation and geotechnics, London, pp.
˜ indicating change 405–412.
 interface friction angle, load transfer ratio k/G, also Chandler, B. C. (1998). My Thuan Bridge: update on bored pile
indicating increment foundations. Proceedings of the Australasian bridge conference,
load transfer parameter Sydney.
non-dimensional pile compressibility parameter Chow, F. C. (1997). Investigations in the behaviour of displacement
º shaft compressibility inverse length, parameter in radial piles for offshore foundations. PhD thesis, Imperial College,
consolidation model London.
 parameter in radial consolidation model and radial Chow, F. C., Jardine, R. J., Brucy, F. & Nauroy, J. F. (1998). Effects
stress degradation of time on capacity of pipe piles in dense marine sand. J.

Poisson’s ratio (subscripts p for pile, r for raft) Geotech. Geoenviron. Engng Div., ASCE 124, No. 3, 254–264.
 ratio of residual to peak shaft friction, diffraction Clancy, P. & Randolph, M. F. (1996). Simple design tools for piled
parameter raft foundations. Géotechnique 46, No. 2, 313–328.
 mathematical constant Cooke, R. W. (1986). Piled raft foundations on stiff clays: a
r area ratio contribution to design philosophy. Géotechnique 36, No. 2,
 stress (subscripts p for vertical yield stress, rc for radial 169–203.
after consolidation, rf for radial at failure, ri for initial Coop, M. R. & Wroth, C. P. (1990). Discussion of M. R. Coop &
radial, v0 for in situ vertical) C. P. Wroth (1989): Field studies of an instrumented model pile
ı shear stress (subscripts peak for peak, res for residual, s in clay. Géotechnique 39, No. 4, 679–696; 40, No. 4, 669–672.
for shaft friction) Coyle, H. M. & Castello, R. R. (1981). New design correlations for
9 soil friction angle piles in sand. J. Geotech. Engng Div., ASCE 197, No. GT7,
ł geometry factor for pile interaction 965–985.
 non-dimensional base stiffness parameter De Beer, E., de Jonghe, A., Carpentier, R. & Wallays, M. (1979).
Analysis of the results of loading tests on displacement piles
penetrating into a very dense sand layer. Proceedings of the
conference on recent developments in the design and construc-
REFERENCES tion of piles, London, pp. 199–211.
Altaee, A., Fellenius, B. H. & Evgin, E. (1992). Axial load transfer Deeks, A. J. & Randolph, M. F. (1995). A simple model for
for piles in sand: I. Tests on an instrumented precast pile. Can. inelastic footing response to transient loading. Int. J. Numer.
Geotech. J. 29, No. 1, 11–20. Anal. Methods Geomech 19, No. 5, 307–329.
Altaee, A., Fellenius, B. H. & Evgin, E. (1993). Axial load transfer De Jong, J. T. & Frost, J. D. (2002). A multisleeve friction
for piles in sand and the critical depth. Can. Geotech. J., 30, attachment for the cone penetrometer. Geotech. Test. J., ASTM
No. 3, 455–463. 25, No. 2, 111–127.
Andersen, K. H. & Jostad, H. P. (2002). Shear strength against De Nicola, A. & Randolph, M. F. (1993). Tensile and compressive
outside wall of suction anchors in clay after installation. Pro- shaft capacity of piles in sand. J. Geotech. Engng Div., ASCE
ceedings of the international conference on offshore and polar 119, No. 12, 1952–1973.
engineering, Kyushu, paper 2002–PCW–02. De Nicola, A. & Randolph, M. F. (1997). The plugging behaviour
API (1993). RP2A: Recommended practice for planning, designing of driven and jacked piles in sand. Géotechnique 47, No. 4,
and constructing fixed offshore platforms. Washington, DC: 841–856.
American Petroleum Institute. De Nicola, A. & Randolph, M. F. (1999). Centrifuge modelling of
Australian Standards (1995). Piling – Design and installation, pipe piles in sand under axial loads. Géotechnique 49, No. 3,
AS2159-1995. Sydney: Standards Australia. 295–318.
Baligh, M. M. (1985). Strain path method. J. Soil Mech. Found. England, M. & Fleming, W. G. K. (1994). Review of foundation
Div., ASCE 111, No. 9, 1180–1136. testing methods and procedures. Proc. ICE Geotech. Engng 107,
Baligh, M. M. (1986). Undrained deep penetration. Géotechnique No. 3, 135–142.
36, No. 4, 471–485; 487–501. Fahey, M. & Lee Goh, A. (1995). A comparison of pressuremeter
Banerjee, P. K. & Butterfield, R. (1981). Boundary element method and piezocone methods of determining the coefficient of con-
in engineering science. New York: McGraw-Hill. solidation, Proc. 4th Int. Symp. on the Pressuremeter and Its
Basile, F. (1999). Non-linear analysis of pile groups. Proc. ICE New Avenues, Quebec, 153–160.
Geotech. Engng 137, 105–115. Fioravante, V. (2002). On the shaft friction modelling of non-
BCP Committee (1971). Field tests on pipe piles in sand. Soils displacement piles in sand. Soils Found. 42, No. 2, 23–33.
Found. 11, No. 2, 29–49. Fioravante, V., Ghionna, V. N., Jamiolkowski, M. & Sarri, H.
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 873
(1999). Shaft friction modelling of non-displacement piles in jacked pipe piles in sand. J. Geotech. Geoenviron. Engng, ASCE
sand. Proceedings of the international conference on analysis, 127, No. 6, 473–480.
design, construction and testing of deep foundations, Austin, Lehane, B. M. & Jardine, R. J. (1994). Displacement–pile behav-
TX. iour in a soft marine clay. Can. Geotech. J. 31, No. 2, 181–191.
Fleming, W. G. K. (1992). A new method for single pile settlement Lehane, B. M. & Randolph, M. F. (2002). Evaluation of a minimum
prediction and analysis. Géotechnique 42, No. 3, 411–425. base resistance for driven pipe piles in siliceous sand. J.
Fleming, W. G. K. & Thorburn, S. (1983). Recent piling advances: Geotech. Geoenviron. Engng Div., ASCE 128, No. 3, 198–205.
state of the art. Proceedings of the conference on recent ad- Lehane, B. M., Jardine, R. J., Bond, A. J. & Frank, R. (1993).
vances in piling and ground treatment, London. Mechanisms of shaft friction in sand from instrumented pile
Franke, E., Lutz, B. & El-Mossallamy, Y. (1994). Measurements tests. J. Geotech. Engng Div., ASCE 119, No. 1, 19–35.
and numerical modelling of high rise building foundations on Lehane, B. M., Jardine, R. J., Bond, A. J. & Chow, F. C. (1994).
Frankfurt clay. Proceedings of the conference on vertical and The development of shaft friction on displacement piles in clay.
horizontal deformations of foundations an embankments, Texas, Proc. 13th Int. Conf. Soil Mech. Found. Engng, New Delhi 2,
ASCE Geotechnical Special Publication No. 40, Vol. 2, pp. 473–476.
1325–1336. Litkouhi, S. & Poskitt, T. J. (1980). Damping constant for pile
Gibson, R. E. & Anderson, W.F. (1961). In situ measurement of soil driveability calculations. Géotechnique 30, No. 1, 77–86.
properties with the pressuremeter, Civ. Engng Public Works Rev. Liyanapathirana, D. S., Deeks, A. J. & Randolph, M. F. (2001).
56, 615–618. Numerical modelling of the driving response of thin-walled
Gregersen, O. S., Aas, G. & Dibagio, E. (1973). Load tests on open-ended piles. Int. J. Numer. Anal. Methods Geomech. 25,
friction piles in loose sand. Proc. 8th Int. Conf. Soil Mech. No. 9, 933–953.
Found. Engng, Moscow 2, 109–117. Maiorano, R. M. S., Viggiani, C. & Randolph, M. F. (1996).
Guo, W. D. & Randolph, M. F. (1997). Vertically loaded piles in Residual stress system arising from different methods of pile
non-homogeneous media. Int. J. Numer. Anal. Methods Geo- installation. Proc. 5th Int. Conf. on Application of Stress-Wave
mech., 21, No. 8, 507–532. Theory to Piles, Orlando, 518–528.
Heerema, E. P. & de Jong, A. (1979). An advanced wave equation Mandolini, A. (2003). Design of piled raft foundations: practice and
computer program which simulates dynamic pile plugging development. Proc. 4th Int. Sem. on Deep Foundations on Bored
through a coupled mass-spring system. Proceedings of the and Auger Piles, Ghent, 59–80.
International conference on numerical methods in offshore Mandolini, A. & Viggiani, C. (1997). Settlement of piled founda-
piling, London, pp. 37–42. tions. Géotechnique 47, No. 4, 791–816.
Hight, D. W., Lawrence, D. M., Farquhar, G. B., Milligan, G. W., Mayne, P. W. (2003). Class ‘A’ footing response prediction from
Gue, S. S. & Potts, D. M. (1996). Evidence for scale effects in seismic cone tests. Proc. 3rd Int. Symp. on Deformation Charac-
the bearing capacity of open-ended piles in sand. Proc. 28th teristics of Geomaterials, Lyon 1, 883–888.
Annual Offshore Technology Conf., Houston, 181–192. Mayne, P. W. & Kulhawy, F. H. (1982). K0 –OCR relationships in
Horikoshi, K. & Randolph, M. F. (1997). On the definition of raft- soils. J. Geotech. Engng Div., ASCE 108, No. GT6, 851–872.
soil stiffness ratio. Géotechnique 47, No. 5, 1055–1061. Mayne, P. W. & Poulos, H. G. (2001). Closure to ‘Approximate
Horikoshi, K. & Randolph, M. F. (1998). Optimum design of piled displacement influence factors for elastic shallow foundations’.
rafts. Géotechnique 48, No. 3, 301–317. J. Geotech. Geoenviron. Engng Div., ASCE 127, No. 1, 100–
Ishihara, K., Saito, A., Shimmi, Y., Miura, Y. & Tominaga, M. 102.
(1977). Blast furnace foundations in Japan. Proc. 9th Int. Conf. McCammon, N. R. & Golder, H. Q. (1970). Some loading tests on
Soil Mech. Found. Engng, Tokyo Case history volume, 157–236. long pipe piles. Géotechnique 20, No. 2, 171–184.
Jardine, R. J. & Chow, F. C. (1996). New design methods for Mesri, G., Rokhsar, A. & Bohor, B. F. (1975). Composition and
offshore piles, MTD Publication 96/103. London: Marine Tech- compressibility of typical samples of Mexico Clay. Géotechni-
nology Directorate. que 25, No. 3, 527–554.
Karlsrud, K. (1999). Lessons learned from instrumented pile load Meyerhof, G. G. (1976). Bearing capacity and settlement of pile
tests in clay. Proceedings of the international conference on foundations. J. Geotech. Engng Div., ASCE 102. No. GT3, 197–228.
analysis, design, construction and testing of deep foundations, Meyerhof, G. G. & Valsangkar, A. J. (1977). Bearing capacity of
Austin, TX. piles in layered soils. Proc. 8th Int. Conf. Soil Mech. Found.
Katzenbach, R., Arslan, U. & Moormann, C. (2000). Piled raft Engng, Moscow 1, 645–650.
foundation projects in Germany. In Design applications of raft Mylonakis, G. (2001). Winkler modulus for axially loaded piles.
foundations (ed. J. A. Hemsley), pp. 323–391, London: Thomas Géotechnique 51, No. 5, 455–461.
Telford. Mylonakis, G. & Gazetas, G. (1998). Settlement and additional
Kirby, R. C. & Esrig, M. I. (1979). Further development of a internal forces of grouped piles in layered soil. Géotechnique
general effective stress method for prediction of axial capacity 48, No. 1, 55–72.
for driven piles in clay. Proceedings of the conference on recent Novak, M. (1977). Vertical vibration of floating piles. J. Engng
developments in the design and construction of piles, London, Mech. Div., ASCE 103, No. EM1, 153–168.
pp. 335–344. O’Neill, M. W. (2001). Side resistance in piles and drilled shafts. J.
Kolk, H. J. & van der Velde, E. (1996). A reliable method to Geotech. Geoenviron. Engng Div., ASCE 127, No. 1, 1–16.
determine friction capacity of piles driven into clays. Proc. Osterberg, J. (1989). New device for load testing driven piles and
Offshore Technology Conf., Houston, Paper OTC 7993. drilled shafts separates friction and end-bearing. Proceedings of
Kulhawy, F. H. (1984). Limiting tip and side resistance: fact or the international conference on piling and deep foundations,
fallacy? In Analysis and design of pile foundations (ed. J. R. London, Vol. 1, pp. 421–427.
Meyer), pp. 80–98. New York: ASCE. Paik, K. H. & Lee, S. R. (1993). Behaviour of soil plugs in open-
Kusakabe, O., Matsumoto, T., Sandanbata, I., Kosuge, S. & ended model piles driven into sands. Mar. Georesources Geo-
Nishimura, S. (1989). Report on questionnaire: predictions of technol. 11, 353–373.
bearing capacity and driveability of piles. Proc. 12th Int. Conf. Paik, K. H., Salgado, R., Lee, J. & Kim, B. (2003). Behaviour of
Soil Mech. Found. Engng, Rio de Janeiro 5, 2957–2963. open and closed-ended piles driven into sand. J. Geotech.
Lee, J. H. & Salgado, R. (1999). Determination of pile base Geoenviron. Engng Div., ASCE 129, No. 4, 296–306.
resistance in sands. J. Geotech. Geoenviron. Engng, ASCE 125, Poulos, H. G. (1968). Analysis of settlement of pile groups.
No. 8, 673–683. Géotechnique 18, No. 3, 449–471.
Lee, S. L., Chow, Y. K., Karunaratne, G. P. & Wong, K. Y. (1988). Poulos, H. G. (1987). Analysis of residual effects in piles. J.
Rational wave equation model for pile-driving analysis. J. Geo- Geotech. Engng Div., ASCE 113, No. 3, 216–219.
tech. Engng, ASCE 114, No. 3, 306–325. Poulos H. G. (1989). Pile behaviour: theory and application.
Lehane, B. M. (1992). Experimental investigations of pile behaviour Géotechnique 39, No. 3, 365–415.
using instrumented field piles. PhD thesis, Imperial College, Poulos, H. G. (1994). An approximate numerical analysis of pile-
London. raft interaction. Int. J. Numer. Anal. Methods Geomech. 18,
Lehane, B. M. & Gavin, K. G. (2001). The base resistance of 73–92.
874 RANDOLPH
Poulos, H. G. (1998). Pile testing: from the designer’s viewpoint. Vesic, A. S. (1970). Tests on instrumented piles, Ogeechee River
Proc. 2nd Int. Statnamic Sem., Tokyo, 3–21. site. J. Soil Mech. Found. Div., ASCE 96, No. SM2, 561–584.
Poulos, H. G. (2001). Piled-raft foundation: design and applications. Viggiani, C. (2001). Analysis and design of piled foundations: First
Géotechnique 51, No. 2, 95–113. Arrigo Croce Lecture. Riv. Ital. di Geotecnica 35, No. 1, 47–75.
Poulos, H. G. & Davis, E. H. (1980). Pile foundation analysis and Vijayvergiya, V. N. & Focht, J. A. (1972). A new way to predict
design. New York: John Wiley & Sons. capacity of piles in clay. Proc. 4th Annual Offshore Technol.
Prakoso, W. A. & Kulhawy, F. H. (2001). Contribution to piled raft Conf., Houston 2, 865–874.
optimum design. J. Geotech. Geoenviron. Engng, ASCE 127, White, D. J. (2003). Field measurements of CPT and pile base
No. 1, 17–24. resistance in sand, Technical Report CUED/D-SOILS/TR327.
Randolph, M. F. (1983). Design considerations for offshore piles. Cambridge University Engineering Dept.
Proceedings of the conference on geotechnical practice in off- White, D. J. & Bolton, M. D. (2002). Observing friction fatigue on
shore engineering, Austin, pp. 422–439. a jacked pile. In Centrifuge and constitutive modelling: two
Randolph, M. F. (1987). Modelling of the soil plug response during extremes (ed. S. M. Springman), pp. 347–354. Rotterdam: Swets
pile driving. Proc. 9th SE Asian Geotech. Conf., Bangkok 2, & Zeitlinger.
6.1–6.14. Whittle, A. J. (1992). Assessment of an effective stress analysis for
Randolph, M. F. (1990). Analysis of the dynamics of pile driving. predicting the performance of driven piles in clays. Proceedings
In Developments in soil mechanics – IV: Advanced geotechnical of the conference on offshore site investigation and foundation
analyses (eds P. K. Banerjee and R. Butterfield). Elsevier behaviour, London, Vol. 28, pp. 607–643.
Applied Science, pp. 223–272.
Randolph, M. F. (1994). Design methods for pile groups and piled VOTE OF THANKS
rafts. Proc. 13th Int. Conf. Soil Mech. Found. Engng, New Delhi HUGH ST JOHN, Director, Geotechnical Consulting Group
5, 61–82.
Randolph, M. F. (2000). Pile–soil interaction for dynamic and static
loading. Proc. 6th Int. Conf. on Application of Stress-Wave When I contacted Mark to ask him to write his own
Theory to Piles, Sao Paulo Appendix, 3–11. introduction (which at the time I thought I was going to
Randolph, M. F. (2003). PIGLET: Analysis and design of pile
do), he said that he would prefer me not to be sycophantic
groups. Users’ Manual, Version 4-2. Perth.
Randolph, M. F. & Deeks, A. J. (1992). Dynamic and static soil but to refer to his return from exile after 16 years. I ended
models for axial pile response dynamics. Proc. 4th Int. Conf. on up taking responsibility for the vote of thanks, which gives
Application of Stress-Wave Theory to Piles, The Hague, 1–14. me a lot more license, because he wouldn’t write that
Randolph, M. F. & Murphy, B. S. (1985). Shaft capacity of driven either.
piles in clay, Proc. 17th Ann. Offshore Technol. Conf., Houston, 1, I would like to start off with a complaint about his
371–378. Australian accent. This was supposed to be an overseas
Randolph, M. F. & Wroth, C. P. (1978). Analysis of deformation of lecture, Mark. You made no effort to make it sound like
vertically loaded piles. J. Geotech. Engng. Div., ASCE 104, No. one.
GT12, 1465–1488. I first met Mark around 30 years ago when he turned up
Randolph, M. F. & Wroth, C. P. (1979). An analytical solution for
the consolidation around a driven pile. Int. J. Numer. Anal.
at BRE, the bright young graduate fresh from Oxford who
Methods Geomech. 3, No. 3, 217–229. seemed a little bewildered about what he was supposed to
Randolph, M. F., Carter, J. P. & Wroth, C. P. (1979). Driven piles in be doing. I had preceded him by a few years but was
clay: the effects of installation and subsequent consolidation. immediately aware of the presence of a superior intellect. I
Géotechnique 29, No. 4, 361–393. gave up competing after we both decided to go in for the
Randolph, M. F., Leong, E. C. & Houlsby, G. T. (1991). One- Cooling Prize. At least I am a quick learner in some
dimensional analysis of soil plugs in pipe piles. Géotechnique respects. I think that these early years were very formative
41, No. 4, 587–598. for him. He found himself in a situation which suited him; a
Randolph, M. F., Dolwin, J. & Beck, R. D. (1994). Design of driven group of people looking for theory to fit their excellent field
piles in sand. Géotechnique 44, No. 3, 427–448.
data, and a leadership which encouraged self development.
Rausche, F., Goble, G. G. & Likins, G. E. (1985). Dynamic
determination of pile capacity. J. Geotech. Engng Div., ASCE Mark soon found his problems to solve and after a short
111, 367–383. while a subject that he could develop as his own, the
Reul, O. & Randolph, M. F. (2004). Design strategies for piled rafts behaviour of piles. Tonight we have seen how this has
subjected to non-uniform vertical loading. J. Geotech. Geoenvir- blossomed, initially under the mentorship of John Burland
on. Engng Div, ASCE 130. (1), (in press). and the late Peter Wroth, of course, two former Rankine
Russo, G. (1998). Numerical analysis of piled rafts. Int. J. Anal. Lecturers. I think that both of them instilled in him what he
Numer. Methods Geomech. 22, No. 6, 477–493. has so ably demonstrated tonight, the importance of the
Semple, R. M. & Rigden, W. J. (1984). Shaft capacity of driven ‘why?’ when deciding on the ‘how?, but also the art of
piles in clay, Proceedings of the symposium on analysis and distilling what is a very complex problem through a series
design of pile foundations, San Francisco, pp. 59–79.
Shioi, Y., Yoshida, O., Meta, T. & Homma, M. (1992). Estimation
of logical steps, to something that can be understood.
of bearing capacity of steel pipe pile by static loading test and Not once this evening have we seen an equation that is
stress-wave theory. Proc. 4th Int. Conf. on Application of Stress- more than half a line long. Not once have we seen a formula
Wave Theory to Piles, The Hauge, 325–330. with a multitude of variables. Mark’s hallmark is that he
Simons, H. A. & Randolph, M. F. (1985). A new approach to one- breaks things down into a series of logical steps and then
dimensional pile driving analysis. Proc. 5th Int. Conf. Numerical uses the building blocks he creates to examine the problem
Methods in Geomechanics, Nagoya 3, 1457–1464. in its entirety. Although this is an academic approach, he
Smith, E. A. L. (1960). Pile driving analysis by the wave equation. uses such models to solve practical problems, and applies
J. Soil Mech., ASCE 86, 35–61. his engineering judgement and observation to challenge the
Teh, C. I. & Houlsby, G. T. (1991). An analytical study of the cone
logic of both other peoples’ theories and the way in which
penetration test in clay. Géotechnique 41, No. 1, 17–34.
Toolan, F. E., Lings, M. L. & Mirza, U. A. (1990). An appraisal of design is carried out. This is a very powerful combination
API RP2A recommendations for determining skin friction of which he has obviously used to good effect in advising a
piles in sand. Proc. 22nd Annual Offshore Technol. Conf., wide range of clients.
Houston, 33–42. Mark expressed some concern to me that in part of his
Vesic, A. S. (1967). A study of bearing capacity of deep founda- lecture he is challenging some of the assumptions made by
tions, Final Report, Project B-189. Atlanta: Georgia Institute of others in the derivation of design methods, and that they
Technology. may take umbrage as they are so close to home. I assured
SCIENCE AND EMPIRICISM IN PILE FOUNDATION DESIGN 875
him, probably unwisely that, as true scientists, trying to between the sky and the size of the rain drops which always
refine their own understanding of their data, they would works.
welcome such a side swipe. But maybe, as a non-academic I Tonight Mark has more than achieved what we have come
haven’t understood how things really work. There are many to expect from a Rankine lecture. He has shown us the fruits
more PhDs to be had from the subject, and I’m sure that of his lifetimes work, challenged us to think further about
even those closest to the subject have seen something here the assumptions that we make when we select a particular
this evening which may stimulate a further thought. design method and shown us ways of doing better. He has
I think that Einstein was right though about psyche of the also demonstrated through some fascinating examples how
Scientist. I hadn’t realised before that even Einstein, like an understanding of the science and thus an appreciation of
everyone else, only presented his best data. . .otherwise he the mechanisms controlling soil–structure interaction can
wouldn’t be sceptical about it. result in finding better solutions to geotechnical problems,
I tried to find a quote to match Mark’s, and decided that I backed up of course by a means of verifying the result. He
should look for something from another adopted son of has been and will continue to be an inspiration to many both
Australia, the late Spike Milligan, who had a theory about within his own highly successful teams and to the geotechni-
the origin of rain. He wrote. . .. cal profession worldwide.
There are holes in the sky where the rain gets in, but they I would like you now to join me in giving our heartfelt
are ever so small. That’s why the rain is thin. thanks to Mark for all the blood, sweat and tears (probably
This is an acute observation, but not backed up by the largely sweat) that he has put in to preparing and delivering
scientific theory. However, it is an empirical relationship the 43rd Rankine Lecture.

Vous aimerez peut-être aussi